Bears of the World: Ecology, Conservation and Management 1108483526, 9781108483520

Bears have fascinated people since ancient times. The relationship between bears and humans dates back thousands of year

1,375 181 114MB

English Pages 500 [437] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Bears of the World: Ecology, Conservation and Management
 1108483526, 9781108483520

  • Commentary
  • Fixed bookmarks and removed duplicate pages

Table of contents :
Cover
Half-page title
Title page
Copyright page
Dedication
Contents
Contributors
Foreward
Acknowledgements
Introduction
Part I : Systematics, Ecology, and Behavior
Chapter 1: Systematics, Evolution, and Genetics of Bears
Chapter 2: Mating Strategies
Chapter 3: Interspecific Interactions between Brown Bears, Ungulates, and Other Large Carnivores
Chapter 4: Adaptations and Competitive Interactions of Tropical Asian Bear Species Define Their Biogeography: Past, Present, and Future
Chapter 5: Remarkable Adaptations of the American Black Bear Help Explain Why it is the Most Common Bear: A Long-Term Study from the Center of its Range
Part II : Species Accounts
Chapter 6: Giant Panda (Ailuropoda melanoleuca)
Chapter 7: Andean Bear (Tremarctos ornatus)
Chapter 8: Sun Bear (Helarctos malayanus)
Chapter 9: Sloth Bear (Melursus ursinus)
Chapter 10: Asiatic Black Bear (Ursus thibetanus)
Chapter 11: American Black Bear (Ursus americanus)
Chapter 12: Brown Bear (Ursus arctos; Eurasia)
Chapter 13: Brown Bear (Ursus arctos; North America)
Chapter 14: Polar Bear (Ursus maritimus)
Part III : Human-Bear Coexistence
Chapter 15: Human–Bear Conflicts at the Beginning of the Twenty-First Century: Patterns, Determinants, and Mitigation Measures
Chapter 16: Principles of Human–Bear Conflict Management in Challenging Environments
Chapter 17: Patterns of Bear Attacks on Humans, Factors Triggering Risky Scenarios, and How to Reduce Them
Chapter 18: Effects of Human Disturbance on Brown Bear Behavior
Chapter 19: Bears in Human-Modified Landscapes: The Case Studies of the Cantabrian, Apennine, and Pindos Mountains
Part IV : Conservation and Manangement
Chapter 20: Conservation and Management of Bears
Chapter 21: How Is Climate Change Affecting Polar Bears and Giant Pandas?
Chapter 22: Managing for Interpopulation Connectivity of the World’s Bear Species
Chapter 23: Ex Situ Conservation of Bears: Roles, Status, and Management
Chapter 24: The Challenge of Brown Bear Management in Hokkaido, Japan
Chapter 25: Potential Ecological Corridors for Remnant Asiatic Black Bear Populations and its Subpopulations Linked to Management Units in Japan
Chapter 26: Captive Bears in Asia: Implications for Animal Welfare and Conservation
Chapter 27: Human Dimensions of Asiatic Black Bear Conflicts and Management in Japan
Chapter 28: Ecological and Social Dimensions of Sloth Bear Conservation in Sri Lanka
Index
Untitled
Plates

Citation preview

Bears of the World Ecology, Conservation and Management Bears have fascinated people since ancient times. The relationship between bears and humans dates back thousands of years, during which time we have also competed with bears for shelter and food. In modern times, bears have come under pressure through encroachment on their habitats, climate change, and illegal trade in their body parts, including the Asian bear bile market. The IUCN lists six bears as Vulnerable or Endangered, and even the Least Concern species, such as the brown bear, are at risk of extirpation in certain countries. The poaching and international trade of these most threatened populations are prohibited, but still ongoing. Covering all bear species worldwide, this beautifully illustrated volume brings together the contributions of 200 international bear experts on the ecology, conservation status, and management of the Ursidae family. It reveals the fascinating long history of interactions between humans and bears and the threats affecting these charismatic species. Vincenzo Penteriani is a researcher at the Spanish National Research Council (CSIC), Senior Editor of Animal Conservation, Associate Editor of Ursus, and is leading the Cantabrian Brown Bear Research Group at the Research Unit of Biodiversity (UMIB), Spain. His current research focuses on the ecology and behavior of brown bears in human-modified landscapes, as well as on large carnivore attacks on humans. Mario Melletti is an independent researcher and member of the African Buffalo Initiative Group, and the Wild Pig Specialist Group, which are part of the International Union for the Conservation of Nature (IUCN SSC). He has contributed to several projects and surveys on both mammals and birds in Africa and Europe. He has edited two books with Cambridge University Press: Ecology, Evolution and Behaviour of Wild Cattle (2014) and Ecology, Conservation and Management of Wild Pigs and Peccaries (2017).

Bears of the World Ecology, Conservation and Management Edited by

Vincenzo Penteriani Spanish National Research Council (CSIC)

Mario Melletti WPSG (Wild Pig Specialist Group), IUCN SSC

University Printing House, Cambridge CB2 8BS, United Kingdom One Liberty Plaza, 20th Floor, New York, NY 10006, USA 477 Williamstown Road, Port Melbourne, VIC 3207, Australia 314–321, 3rd Floor, Plot 3, Splendor Forum, Jasola District Centre, New Delhi – 110025, India 79 Anson Road, #06–04/06, Singapore 079906 Cambridge University Press is part of the University of Cambridge. It furthers the University’s mission by disseminating knowledge in the pursuit of education, learning, and research at the highest international levels of excellence. www.cambridge.org Information on this title: www.cambridge.org/9781108483520 DOI: 10.1017/9781108692571 © Cambridge University Press 2021 This publication is in copyright. Subject to statutory exception and to the provisions of relevant collective licensing agreements, no reproduction of any part may take place without the written permission of Cambridge University Press. First published 2021 Printed in the United Kingdom by TJ Books Ltd, Padstow Cornwall A catalogue record for this publication is available from the British Library. ISBN 978-1-108-48352-0 Hardback Cambridge University Press has no responsibility for the persistence or accuracy of URLs for external or third-party internet websites referred to in this publication and does not guarantee that any content on such websites is, or will remain, accurate or appropriate.

To those Apennine brown bears who, in the eighties, showed a young boy how to live following the rhythm of the seasons; to my father Stefano, my daughter Giulia and María del Mar, who have always understood my long absences from bear den exit to the beginning of hibernation. Vincenzo Penteriani To my mother Carla, for the amazing woman she is, whose support, advice, and encouragement throughout my life helped me to pursue my dreams! To my family, Marco, Lucia, Luca, Chiara, Jacopo, Diletta, Alessandro, Kuma, India, Rhum, Kelly, and Kira for your help and patience during my absences. Mario Melletti

Contents List of Contributors xi Foreword by Tim Clutton-Brock Acknowledgments xxi

Introduction

xix

Part I – Systematics, Ecology, and Behavior 1.

Systematics, Evolution, and Genetics of Bears Andrew C. Kitchener, Eva Bellemain, Xiang Ding, Alexander Kopatz, Verena E. Kutschera, Valentina Salomashkina, Manuel Ruiz-García, Tabitha Graves, Yiling Hou, Lars Werdelin, and Axel Janke

2.

Mating Strategies 21 Sam M.J.G. Steyaert, Andreas Zedrosser, Ronald R. Swaisgood, Eva Filipczykova, Brian Crudge, Trishna Dutta, Sandeep Sharma, Shyamala Ratnayeke, Shinsuke Koike, Martin Leclerc, Andrew E. Derocher, Melanie Clapham, Thomas Spady, Bruce McLellan, Andrés Ordiz, Alberto FernándezGil, Miguel Delibes, and Jon E. Swenson

3.

4.

5.

Part II – Species Accounts

1

Interspecific Interactions between Brown Bears, Ungulates, and Other Large Carnivores 36 Andrés Ordiz, Miha Krofel, Cyril Milleret, Ivan Seryodkin, Aimee Tallian, Ole-Gunnar Støen, Therese Ramberg Sivertsen, Jonas Kindberg, Petter Wabakken, Håkan Sand, and Jon E. Swenson Adaptations and Competitive Interactions of Tropical Asian Bear Species Define Their Biogeography: Past, Present, and Future 45 Robert Steinmetz, David L. Garshelis, and Anwaruddin Choudhury Remarkable Adaptations of the American Black Bear Help Explain Why it is the Most Common Bear: A Long-Term Study from the Center of its Range 53 David L. Garshelis, Karen V. Noyce, Mark A. Ditmer, Pamela L. Coy, Andrew N. Tri, Timothy G. Laske, and Paul A. Iaizzo

6.

Giant Panda (Ailuropoda melanoleuca) 63 Ronald R. Swaisgood, William M. McShea, David Wildt, Vanessa Hull, Jindong Zhang, Megan A. Owen, Zejun Zhang, Zachary Dvornicky-Raymond, Marc Valitutto, Dihua Li, Zhang Hemin, Jenny Santiestevan, and Fuwen Wei

7.

Andean Bear (Tremarctos ornatus) 78 Ximena Velez-Liendo, David Jackson, Manuel Ruiz-García, Armando Castellanos, Santiago Espinosa, and Andrés Laguna

8.

Sun Bear (Helarctos malayanus) 88 Lorraine Scotson, Cheryl Frederick, Kirsty Officer, and Wai-Ming Wong

9.

Sloth Bear (Melursus ursinus) 99 Harendra Singh Bargali, Karine E. Pigeon, Naim Akhtar, Thomas Sharp, and Kajal K. Jadav

3

10. Asiatic Black Bear (Ursus thibetanus) Chinatsu Kozakai, Ivan Seryodkin, Karine E. Pigeon, Koji Yamazaki, Sangay Wangchuk, Shinsuke Koike, Toshio Tsubota, and Yonten Jamtsho

110

11. American Black Bear (Ursus americanus) Joseph D. Clark, Jon P. Beckmann, Mark S. Boyce, Bruce D. Leopold, Anne E. Loosen, and Michael R. Pelton

122

12. Brown Bear (Ursus arctos; Eurasia) 139 Jon E. Swenson, Hüseyin Ambarlı, Jon M. Arnemo, Leonid Baskin, Paolo Ciucci, Pjotr I. Danilov, Miguel Delibes, Marcus

vii

Contents

Elfström, Alina L. Evans, Claudio Groff, Anne G. Hertel, Djuro Huber, Klemen Jerina, Alexandros A. Karamanlidis, Jonas Kindberg, Ilpo Kojola, Miha Krofel, Josip Kusak, Tsutomu Mano, Mario Melletti, Yorgos Mertzanis, Andrés Ordiz, Santiago Palazón, Jamshid Parchizadeh, Vincenzo Penteriani, Pierre-Yves Quenette, Agnieszka Sergiel, Nuria Selva, Ivan Seryodkin, Michaela Skuban, Sam M.J.G. Steyaert, Ole-Gunnar Støen, Konstantin F. Tirronen, and Andreas Zedrosser 13. Brown Bear (Ursus arctos; North America) Mark A. Haroldson, Melanie Clapham, Cecily C. Costello, Kerry A. Gunther, Katherine C. Kendall, Sterling D. Miller, Karine E. Pigeon, Michael F. Proctor, Karyn D. Rode, Christopher Servheen, Gordon B. Stenhouse, and Frank T. van Manen

Yamazaki, Harendra Singh Bargali, Nishith Dharaiya, Ashish Kumar Jangid, Ravi Kumar Sharma, Babu Ram Lamichhane, Shyamala Ratnayeke, Ivan Seryodkin, Himanshu Shekhar Palei, Ashok Subedi, Hüseyin Ambarlı, José María Fedriani, Pedro José Garrote, Klemen Jerina, Ilpo Kojola, Miha Krofel, Prakash Mardaraj, Mario Melletti, Andrés Ordiz, Paolo Pedrini, Eloy Revilla, Luca Francesco Russo, Veronica Sahlén, Christopher Servheen, Ole-Gunnar Støen, Jon E. Swenson, and Tom Smith 162

14. Polar Bear (Ursus maritimus) 196 Karyn D. Rode, Martyn Obbard, Stanislav E. Belikov, Andrew E. Derocher, George M. Durner, Gregory W. Thiemann, Morten Tryland, Robert J. Letcher, Randi Meyerson, Christian Sonne, Bjørn M. Jenssen, Rune Dietz, and Dag Vongraven

Part III – Human–Bear Coexistence 15. Human–Bear Conflicts at the Beginning of the Twenty-First Century: Patterns, Determinants, and Mitigation Measures 213 Miha Krofel, Marcus Elfström, Hüseyin Ambarlı, Giulia Bombieri, Enrique González-Bernardo, Klemen Jerina, Andrés Laguna, Vincenzo Penteriani, James P. Phillips, Nuria Selva, Seth M. Wilson, Alejandra Zarzo-Arias, Claudio Groff, Djuro Huber, Alexandros A. Karamanlidis, Yorgos Mertzanis, Eloy Revilla, and Carlos Bautista 16. Principles of Human–Bear Conflict Management in Challenging Environments 227 Özgün Emre Can 17. Patterns of Bear Attacks on Humans, Factors Triggering Risky Scenarios, and How to Reduce Them 239 Vincenzo Penteriani, Giulia Bombieri, María del Mar Delgado, Thomas Sharp, Koji

viii

18. Effects of Human Disturbance on Brown Bear Behavior 250 Ole-Gunnar Støen, Andrés Ordiz, Marcus Elfström, Anne G. Hertel, Veronica Sahlén, Jonas Kindberg, and Jon E. Swenson 19. Bears in Human-Modified Landscapes: The Case Studies of the Cantabrian, Apennine, and Pindos Mountains 260 Vincenzo Penteriani, Alexandros A. Karamanlidis, Andrés Ordiz, Paolo Ciucci, Luigi Boitani, Giorgio Bertorelle, Alejandra Zarzo-Arias, Giulia Bombieri, Enrique González-Bernardo, Paola Morini, Francesco Pinchera, Néstor Fernández, María C. Mateo-Sánchez, Eloy Revilla, Miguel de Gabriel Hernando, Yorgos Mertzanis, and Mario Melletti

Part IV – Conservation and Management 20. Conservation and Management of Bears Christopher Servheen, Hüseyin Ambarlı, Harendra Singh Bargali, Stewart W. Breck, Neil D’Cruze, Claudio Groff, Gabriella M. Fredriksson, Michael L. Gibeau, Isaac Goldstein Aizman, Djuro Huber, Katherine C. Kendall, Sterling D. Miller, Michael F. Proctor, Murray Rutherford, Lorraine Scotson, and Jon E. Swenson

273

21. How Is Climate Change Affecting Polar Bears and Giant Pandas? 303 Melissa Songer, Todd C. Atwood, David C. Douglas, Qiongyu Huang, Renqiang Li, Nicholas W. Pilfold, Ming Xu, and George M. Durner 22. Managing for Interpopulation Connectivity of the World’s Bear Species 317 Michael F. Proctor, Trishna Dutta, Bruce N. McLellan, Shaenandhoa Garcia Rangel,

Contents

Dave Paetkau, Ronald R. Swaisgood, and Andreas Zedrosser 23. Ex Situ Conservation of Bears: Roles, Status, and Management 338 Lydia Kolter, Agnieszka Sergiel, Djuro Huber, and Scott Silver 24. The Challenge of Brown Bear Management in Hokkaido, Japan 349 Tsutomu Mano, Masami Yamanaka, Hifumi Tsuruga, and Yoshikazu Sato 25. Potential Ecological Corridors for Remnant Asiatic Black Bear Populations and its Subpopulations Linked to Management Units in Japan 356 Tomoko Doko, Wenbo Chen, Reina Uno, Hidetoshi B. Tamate, A.G. Toxopeus, A.K. Skidmore, and Hiromichi Fukui

26. Captive Bears in Asia: Implications for Animal Welfare and Conservation Jan Schmidt-Burbach, Fakhar-i-Abbas, and Neil D’Cruze

364

27. Human Dimensions of Asiatic Black Bear Conflicts and Management in Japan 370 Ryo Sakurai 28. Ecological and Social Dimensions of Sloth Bear Conservation in Sri Lanka 379 Shyamala Ratnayeke and Frank T. van Manen

Index 387 Color plates can be found between pages 202 and 203.

ix

Contributors

Fakhar-i-Abbas Bioresource Research Centre Isaac Goldstein Aizman Wildlife Conservation Society (WCS), New York, USA Naim Akhtar Nature Activity Centre for Environment Education, Ahmedabad, India Hüseyin Ambarlı Department of Wildlife Ecology and Management, Faculty of Forestry, Düzce University, Düzce, Turkey Jon Martin Arnemo Department of Forestry and Wildlife Management, Inland Norway University of Applied Sciences, Elverum, Norway and Department of Wildlife, Fish, and Environmental Studies, Swedish University of Agricultural Sciences, Umeå, Sweden Todd C. Atwood US Geological Survey, Alaska Science Center, Anchorage, AK, USA Harendra Singh Bargali The Corbett Foundation, Uttarakhand, India Leonid Baskin Institute of Ecology and Evolution, Russian Academy of Sciences, Moscow, Russia Carlos Bautista Institute of Nature Conservation, Polish Academy of Sciences, Krakow, Poland Jon P. Beckmann Wildlife Conservation Society, North America Program and University of Nevada, Reno, USA Stanislav E. Belikov All-Russian Research Institute for Environment Protection, Sarov, Russia

Eva Bellemain SPYGEN, Savoie Technolac, Le Bourget du Lac, France Giorgio Bertorelle Department of Life Sciences and Biotechnology, University of Ferrara, Ferrara, Italy Luigi Boitani Department of Biology and Biotechnologies “Charles Darwin”, Rome, Italy Giulia Bombieri Research Unit of Biodiversity, Oviedo University – Campus Mieres, Mieres, Spain Museo delle Scienze, Sezione Zoologia dei Vertebrati, Trento, Italy Mark S. Boyce Department of Biological Sciences, University of Alberta, Edmonton, Canada Stewart W. Breck USDA-National Wildlife Research Center, Department of Fish, Wildlife, & Conservation Biology, Colorado State University, CO, USA Jan Schmidt-Burbach The World Society for the Protection of Animals, London, UK Özgün Emre Can Wildlife Conservation Research Unit (WildCRU), Department of Zoology, University of Oxford, UK Armando Castellanos Andean Bear Foundation, La Isla, Quito, Ecuador TSG (Tapir Specialist Group) IUCN SSC, Gland, Switzerland

xi

List of Contributors

Wenbo Chen Graduate School of Media and Governance, Keio University, Japan

Nishith Dharaiya Department of Life Sciences, Hemchandracharya North Gujarat University, Patan, India

Nature & Science Consulting Co., Kanagawa, Japan

Rune Dietz Faculty of Science and Technology, Department of Bioscience, Arctic Research Center, Aarhus University, Aarhus, Denmark

Anwaruddin Choudhury The Rhino Foundation for Nature in NE India, Guwahati, India Paolo Ciucci Department of Biology and Biotechnologies “Charles Darwin”, Università di Roma “La Sapienza”, Rome, Italy

Xiang Ding Key Laboratory of Southwest China Wildlife Resources Conservation, College of Life Sciences, China West Normal University, Nanchong, China

Melanie Clapham Applied Conservation Science Laboratory, Department of Geography, University of Victoria, Victoria, Canada

Mark A. Ditmer Department of Fisheries, Wildlife and Conservation Biology, University of Minnesota, MN, USA

Joseph D. Clark Southern Appalachian Field Branch, Northern Rocky Mountain Science Center, US Geological Survey, Knoxville, TN, USA

Tomoko Doko Graduate School of Media and Governance, Keio University, Fujisawa, Japan

Cecily C. Costello Research Wildlife Biologist, Montana Fish, Wildlife & Parks, Helena, USA Pamela L. Coy Minnesota Department of Natural Resources, Minnesota, USA Brian Crudge Free The Bears, Luang Prabang, Lao PDR Department of Natural Resources and Environmental Health, University of South-Eastern Norway, Telemark, Norway Neil D’Cruze PanEco, Yayasan Ekosistem Lestari, Sumatran Orangutan Conservation Programme, Pro Natura Foundation, Medan, Indonesia Pjotr I. Danilov Laboratory of Zoology, Institute of Biology, Karelian Research Centre of Russian Academy of Sciences, Petrozavodsk, Russia María del Mar Delgado Research Unit of Biodiversity, Oviedo University – Campus Mieres, Mieres, Spain Miguel Delibes Department of Conservation Biology, Estación Biológica de Doñana, CSIC (Spanish Council of Scientific Research), Seville, Spain Andrew E. Derocher Department of Biological Sciences, University of Alberta, Edmonton, Canada

xii

Nature & Science Consulting Co., Ltd., Yokohama, Japan David C. Douglas US Geological Survey, Alaska Science Center, Juneau, AK, USA George M. Durner US Geological Survey, Alaska Science Center, Anchorage, AK, USA Trishna Dutta Wildlife Sciences, Faculty of Forest Sciences and Forest Ecology, University of Goettingen, Goettingen, Germany Zachary Dvornicky-Raymond Global Health Program, Smithsonian Conservation Biology Institute, Front Royal, VA, USA Marcus Elfström EnviroPlanning AB, Gothenburg, Sweden Norwegian University of Life Sciences, Faculty of Environmental Sciences and Natural Resource Management, Ås, Norway Santiago Espinosa Facultad de Ciencias UASLP, San Luis Potosí, México Alina L. Evans Department of Forestry and Wildlife Management, Inland Norway University of Applied Sciences, Elverum, Norway José María Fedriani Centro de Investigaciones sobre Desertificación (CIDE-CSIC), Moncada (Valencia), Spain Centre for Applied Ecology “Prof. Baeta Neves”/InBIO, Institute of Agronomy, University of Lisbon, Lisbon, Portugal

List of Contributors

Néstor Fernández German Centre for Integrative Biodiversity Research (iDiv), Halle-Jena-Leipzig, Germany Alberto Fernández-Gil Department of Conservation Biology, Estación Biológica de Doñana, CSIC (Spanish Council of Scientific Research), Seville, Spain Eva Filipczykova School of Health, Medical and Applied Sciences, Gladstone, Australia Czech University of Life Sciences Prague, Faculty of Environmental Sciences, Prague, Czech Republic Cheryl Frederick School of Biodiversity Conservation, Unity College, Unity, ME, USA Gabriella M. Fredriksson Institute for Biodiversity and Ecosystem Dynamics/Zoological Museum, University of Amsterdam, Amsterdam, The Netherlands Hiromichi Fukui International Digital Earth Applied Science Research Center, Chubu Institute for Advanced Studies, Chubu University, Kasugai, Japan Pedro José Garrote Centre for Applied Ecology “Prof. Baeta Neves”/InBIO, Institute of Agronomy, University of Lisbon, Lisbon, Portugal David L. Garshelis Minnesota Department of Natural Resources, MN, USA Michael L. Gibeau Mountain National Parks, Parks Canada, Lake Louise, Canada

Mark A. Haroldson US Geological Survey, Northern Rocky Mountain Science Center, Interagency Grizzly Bear Study Team, MT, USA Zhang Hemin China Conservation and Research Center for the Giant Panda, Chengdu Dujiangyan, China Miguel de Gabriel Hernando Department of Biodiversity and Environmental Management, Faculty of Biological and Environmental Sciences, Universidad de León, León, Spain Anne G. Hertel Senckenberg Biodiversity and Climate Research Centre, Frankfurt (Main), Germany Yiling Hou Key Laboratory of Southwest China Wildlife Resources Conservation (Ministry of Education), College of Life Sciences, China West Normal University, Nanchong, China Qiongyu Huang Smithsonian Conservation Biology Institute, Front Royal, VA, USA Djuro Huber Institute of Nature Conservation of Polish Academy of Sciences, Adama Mickiewicza, Krakow, Poland Vanessa Hull Department of Wildlife Ecology and Conservation, University of Florida, Gainesville, FL, USA Paul A. Iaizzo Department of Surgery, University of Minnesota, Saint Paul, MN, USA David Jackson Andean Bear Foundation, La Isla, Quito, Ecuador

Enrique González-Bernardo Research Unit of Biodiversity, Oviedo University – Campus Mieres, Mieres, Spain

Kajal K. Jadav School of Wildlife Forensic and Health, Nanaji Deshmukh Veterinary Science, Jabalpur, India

Tabitha Graves US Geological Survey, Northern Rocky Mountain Science Center, West Glacier, MT, USA

Yonten Jamtsho Wangchuck Centennial National Park, Department of Forest and Park Services, Ministry of Agriculture and Forest, Bhutan

Claudio Groff Servizio Foreste e Faune, Settore Grandi Carnivori Via G. B. Trener, Trento, Italy Kerry A. Gunther Bear Management Office, Yellowstone National Park, WY, USA

Ashish Kumar Jangid Tiger Cell, Wildlife Institute of India, Dehradun, India Axel Janke Department of Evolutionary Biology, Evolutionary Biology Centre, Uppsala University, Uppsala, Sweden

xiii

List of Contributors

Bjørn M. Jenssen Department of Biology, Norwegian University of Science and Technology, Trondheim, Norway

Andrés Laguna Direction of Research Department, Big Mammals Conservation Company, Ecuador

Klemen Jerina Department of Forestry and Renewable Forest Resources, Biotechnical Faculty, University of Ljubljana, Ljubljana, Slovenia

Babu Ram Lamichhane NTNC – Biodiversity Conservation Center, Sauraha, Nepal

Alexandros A. Karamanlidis ARCTUROS – Civil Society for the Protection and Management of Wildlife and the Natural Environment, Aetos, Greece

Institute of Cultural Anthropology and Development Sociology, Leiden University, Leiden, The Netherlands

Katherine C. Kendall US Geological Survey, Reston, VA, USA

Timothy G. Laske Department of Surgery, University of Minnesota, Saint Paul, MN, USA

Jonas Kindberg Norwegian Institute for Nature Research, Trondheim, Norway and Department of Wildlife, Fish, and Environmental Studies, Swedish University of Agricultural Sciences, Umea, Sweden

Martin Leclerc Applied Conservation Science Laboratory, Department of Geography, University of Victoria, Victoria, Canada

Andrew C. Kitchener Department of Natural Sciences, National Museums Scotland, Edinburgh, UK

Bruce D. Leopold Wildlife Ecology, Michigan State University, East Lansing, MI, USA

Shinsuke Koike Institute of Agriculture, Tokyo University of Agriculture and Technology, Tokyo, Japan

Robert J. Letcher Environment and Climate Change Canada, Wildlife and Landscape Science Directorate, Ecotoxicology and Wildlife Health Division, National Wildlife Research Centre, Carleton University, Carleton, Ottawa, Canada

Ilpo Kojola Natural Resources Institute Finland (Luke), Rovaniemi, Finland Lydia Kolter IUCN SSC Bear Specialist Group, Cologne, Germany Alexander Kopatz Norwegian Institute for Nature Research, Trondheim, Norway Chinatsu Kozakai Central Region Agricultural Research Center, National Agriculture and Food Research Organization (NARO), Ibaraki, Japan Miha Krofel Department of Forestry and Renewable Forest Resources, Biotechnical Faculty, University of Ljubljana, Ljubljana, Slovenia Josip Kusak Department of Biology, University of Zagreb, Zagreb, Croatia Verena E. Kutschera Science for Life Laboratory, Department of Biochemistry and Biophysics, Stockholm University, Solna, Sweden

xiv

Evolutionary Ecology Group, Faculty of Sciences, University of Antwerp, Antwerp, Belgium

Dihua Li Urban and Regional Ecology and Planning, Landscape Architecture, Landscape Sociology, College of Architecture and Landscape, Peking University, Beijing, China Renqiang Li Key Laboratory of Ecosystem Network Observation and Modeling, Institute of Geographic Sciences and Natural Resources Research, Chinese Academy of Sciences, Beijing, China Anne E. Loosen Department of Biological Sciences, University of Alberta, Edmonton, Canada Frank T. van Manen US Geological Survey, Northern Rocky Mountain Science Center, Interagency Grizzly Bear Study Team, MT, USA Tsutomu Mano Nature Conservation Division, Institute of Environmental Sciences, Environmental and Geological Research Organization, Japan Prakash Mardaraj IUCN/Sloth Bear & Human–Bear Conflict Expert Team, Amity Institute of Forestry & Wildlife, Noida, India

List of Contributors

María C. Mateo-Sánchez ETSI Montes, Forestal y del Medio Natural, Technical University of Madrid, Ciudad Universitaria, Madrid, Spain Bruce McLellan Emeritus Scientist, British Columbia Forest Service, D’Arcy, British Columbia, Canada William M. McShea Conservation Ecology Center, Smithsonian Conservation Biology Institute, National Zoological Park, VA, USA Mario Melletti Wild Pig Specialist Group (WPSG), IUCN SSC, Gland, Switzerland Yorgos Mertzanis “CALLISTO” NGO, Wildlife and Nature Conservation Society Mitropoleos, Thessaloniki, Greece Randi Meyerson Detroit Zoological Society, Detroit, MI, USA Sterling D. Miller Alaska Department of Fish and Game and National Wildlife Federation, Lolo, MT, USA Cyril Milleret Faculty of Environmental Sciences and Natural Resource Management, Norwegian University of Life Sciences, Ås, Norway Paola Morini Sirente Velino Natural Regional Park – Viale XXIV Maggio snc, Rocca di Mezzo (AQ), Italy Karen V. Noyce Minnesota Department of Natural Resources, MN, USA Martyn Obbard Wildlife Research and Monitoring Section, Ontario Ministry of Natural Resources and Forestry, DNA Building, Trent University, Peterborough, Canada Kirsty Officer Free the Bears, Phnom Penh, Cambodia Andrés Ordiz Faculty of Environmental Sciences and Natural Resource Management, Norwegian University of Life Sciences, Ås, Norway Megan A. Owen Recovery Ecology, Institute for Conservation Research, San Diego Zoo Global, Escondido, CA, USA

Dave Paetkau Wildlife Genetics International, Nelson, British Columbia, Canada Santiago Palazón Fauna and Flora Service, Government of Catalonia, Barcelona, Spain Himanshu Shekhar Palei Aranya Foundation, Bhubaneswar, India Jamshid Parchizadeh Global Wildlife Conservation Center, State University of New York College of Environmental Science and Forestry, Syracuse, NY, USA Paolo Pedrini Muse – Museo delle Scienze, Trento, Italy Michael R. Pelton Institute of Agriculture, University of Tennessee, Knoxville, TN, USA Vincenzo Penteriani Spanish National Research Council (CSIC), Research Unit of Biodiversity (UMIB), Mieres, Spain James P. Phillips Oregon State University, Corvallis, OR, USA Karine E. Pigeon Geomatics and Landscape Ecology Laboratory (GLEL), Carleton University, Ottawa, Canada Nicholas W. Pilfold Applied Animal Ecology, Institute for Conservation Research, San Diego Zoo Global, Escondido, CA, USA Francesco Pinchera CISDAM – Via Liberata 1, Rosello (CH), Italy Michael F. Proctor Birchdale Ecological, Kaslo, Canada Pierre-Yves Quenette Office Français de la Biodiversité, Direction de la Recherche et de l’Appui Scientifique, Villeneuve de Rivière, France Shaenandhoa Garcia Rangel United Nations Environment World Conservation Monitoring Program, Cambridge, UK Shyamala Ratnayeke Department of Biological Sciences, Sunway University, No. 5 Jalan University, Selangor, Malaysia Eloy Revilla Pyrenean Institute of Ecology, Zaragoza, Spain

xv

List of Contributors

Karyn D. Rode US Geological Survey, Alaska Science Center, Anchorage, AK, USA

Ivan Seryodkin Laboratory of Animals Ecology and Conservation, Pacific Geographical Institute FEB RAS, Vladivostok, Russia

Manuel Ruiz-García Laboratorio de Genética de Poblaciones Molecular-Biología Evolutiva, Departamento de Biología, Facultad de Ciencias, Pontificia Universidad Javeriana, Bogotá DC, Colombia

Ravi Kumar Sharma Tiger Cell, Wildlife Institute of India, Dehradun, India

Luca Francesco Russo Dipartimento Bioscienze e Territorio, Contrada Fonte Lappone, Pesche, Italy Murray Rutherford School of Resource and Environmental Management, Simon Frazer University, Burnaby, Canada Veronica Sahlén Faculty of Environmental Sciences and Natural Resource Management, Norwegian University of Life Sciences, Ås, Norway Ryo Sakurai College of Policy Science, Ritsumeikan University, Osaka, Japan Valentina Salomashkina A.N. Severtsov Institute of Ecology and Evolution, Russian Academy of Sciences, Moscow, Russia National Research Center for Hematology, Ministry of Healthcare of the Russian Federation, Moscow, Russia

Thomas Sharp Conservation and Research/Wildlife SOS – USA, Salt Lake City, UT, USA Scott Silver National Audubon Society, Constitution Marsh Audubon Center and Sanctuary, New York, USA Therese Ramberg Sivertsen Norwegian Institute for Nature Research, Oslo, Norway A.K. Skidmore Department of Natural Resources, ITC, Faculty of Geo-Information Science and Earth Observation of the University of Twente, Enschede, The Netherlands Michaela Skuban Carpathian Wildlife Society, Zvolen, Slovakia

Håkan Sand Swedish University of Agricultural Sciences, Uppsala, Sweden

Tom Smith Department of Plant and Wildlife Sciences, Brigham Young University, Provo, UT, USA

Jenny Santiestevan Global Health Program, Smithsonian Conservation Biology Institute, Front Royal, VA, USA

Melissa Songer Smithsonian Conservation Biology Institute, Front Royal, VA, USA

Yoshikazu Sato Rakuno Gakuen University, Ebetsu, Hokkaido, Japan

Christian Sonne Arctic Research Center, Department of Bioscience, Faculty of Science and Technology, Aarhus University, Aarhus, Denmark

Lorraine Scotson IUCN Saola Working Group, Vinh University, Vinh, Vietnam Nuria Selva Institute of Nature Conservation, Polish Academy of Sciences, Krakow, Poland Agnieszka Sergiel Institute of Nature Conservation, Polish Academy of Sciences, Krakow, Poland Christopher Servheen Dept. of Ecosystem and Conservation Sciences, W.A. Franke College of Forestry and Conservation, University of Montana Missoula, MT, USA

xvi

Sandeep Sharma Department of Conservation Biology, Johann-FriedrichBlumenbach Institute of Zoology, University of Goettingen, Goettingen, Germany

Thomas Spady California State University San Marcos, San Marcos, CA, USA Robert Steinmetz WWF-Thailand, Bangkok, Thailand Gordon B. Stenhouse Foothills Research Institute, Hinton, Alberta, Canada Sam M.J.G. Steyaert Faculty of Biosciences and Aquaculture, Nord University, Steinkjer, Norway

List of Contributors

Ole-Gunnar Støen Norwegian Institute for Nature Research, Oslo, Norway Faculty of Environmental Sciences and Natural Resource Management, Norwegian University of Life Sciences, Ås, Norway Ashok Subedi National Trust for Nature Conservation, Annapurna Conservation Area Project, Pokhara, Nepal Ronald R. Swaisgood Recovery Ecology, Institute for Conservation Research, San Diego Zoo Global, Escondido, CA, USA Jon E. Swenson Faculty of Environmental Sciences and Natural Resource Management, Norwegian University of Life Sciences, Ås, Norway Aimee Tallian Norwegian Institute for Nature Research, Trondheim, Norway Hidetoshi B. Tamate Department of Biology, Yamagata University, Yamagata, Japan

Reina Uno Institute for Advanced Biosciences, Keio University, Tsuruoka, Japan Marc Valitutto Smithsonian Conservation Biology Institute’s Global Health Program, Front Royal, VA, USA Ximena Velez-Liendo Wildlife Conservation Research Unit, Recanati-Kaplan Centre, Oxford, UK Chester Zoo, Cedar House, Chester, UK Dag Vongraven Norwegian Polar Institute, Tromsø, Norway Petter Wabakken Department of Forestry and Wildlife Management, Inland Norway University of Applied Sciences, Elverum, Norway Sangay Wangchuk Ugyen Wangchuck Institute for Conservation and Environment Research, Bhutan

Gregory W. Thiemann Faculty of Environmental Studies, York University, Toronto, Canada

Fuwen Wei Institute of Zoology, Chinese Academy of Sciences (CAS), Key Laboratory of Animal Ecology and Conservation Biology, Beijing, China

Konstantin F. Tirronen Laboratory of Zoology, Institute of Biology, Karelian Research Centre of Russian Academy of Sciences, Petrozavodsk, Russia

Lars Werdelin Department of Palaeobiology, Swedish Museum of Natural History, Stockholm, Sweden

A.G. Toxopeus Department of Natural Resources, ITC, Faculty of GeoInformation Science and Earth Observation of the University of Twente, Enschede, The Netherlands

David Wildt Center for Species Survival, Smithsonian Conservation Biology Institute, Front Royal, VA, USA

Andrew N. Tri Minnesota Department of Natural Resources, Grand Rapids, MN, USA

Seth M. Wilson W.A. Franke College of Forestry & Conservation – University of Montana, Missoula, MT, USA

Morten Tryland UiT The Arctic University of Norway, Arctic and Marine Biology, Arctic Infection Biology, Tromsø, Norway

Northern Rockies Conservation Cooperative, Jackson, WY, USA

Toshio Tsubota Laboratory of Wildlife Biology and Medicine, Department of Environmental Veterinary Sciences, Graduate School of Veterinary Medicine, Hokkaido University, Sapporo, Japan Hifumi Tsuruga Hokkaido Research Organization, Japan Research Institute of Energy, Environment and Geology, Industrial Technology and Environment Research Department, Sapporo, Hokkaido, Japan

Wai-Ming Wong Panthera, New York, USA Ming Xu Department of Ecology, Evolution, and Natural Resources, Grant F. Walton Center for Remote Sensing & Spatial Analysis, School of Environmental and Biological Sciences, Rutgers University, New Brunswick, NJ, USA Masami Yamanaka Shari Town Museum, Japan

xvii

List of Contributors

Koji Yamazaki Forest Ecology Laboratory, Department of Forest Science, Faculty of Regional Environmental Science, Tokyo University of Agriculture, Tokyo, Japan

Museo delle Scienze, Sezione Zoologia dei Vertebrati, Trento, Italy

Alejandra Zarzo-Arias Research Unit of Biodiversity, Oviedo University – Campus Mieres, Mieres, Spain

Alberto Mattia Perronace Rome, Italy

Andreas Zedrosser Department of Natural Resources and Environmental Health University of South-Eastern Norway, Bø, Norway Jindong Zhang Key Laboratory of Southwest China Wildlife Resources Conservation, China West Normal University, Ministry of Education, Nanchong, China Zejun Zhang China West Normal University, Nanchong, China

Illustrators Giulia Bombieri Research Unit of Biodiversity, Oviedo University – Campus Mieres, Mieres, Spain

xviii

Enrico Chiarelli Rome, Italy

Francesco Rinaldi Sansepolcro, Italy Luciano Toma Dipartimento di Malattie Infettive, Reparto di Malattie Trasmesse da Vettori, Istituto Superiore di Sanità, Rome, Italy

Distribution maps Daniele Baisero Key Biodiversity Area Secretariat, Wildlife Conservation Society, Cambridge, UK

Foreword

Bears of the World brings together for the first time much of what is known of the evolution, ecology, behavior, and conservation status of the eight existing species of bears in a single volume. Its 29 chapters are written mostly by collaborating groups of scientists from multiple institutions and countries and provide biologists and managers concerned with bears with easy access to balanced, up-to-date accounts of different aspects of the biology, ecology, and conservation of bears and a single go-to source of information. The first five chapters deal with general comparative topics – systematics, mating strategies, reproduction, comparative anatomy, and ecology of bears – while the next nine provide focused reviews of the existing knowledge of the biology, ecology, and conservation status of different bear species: giant pandas, Andean bears, sun bears, sloth bears, Asiatic and North American black bears, brown bears (including grizzlies and Kodiak bears), and polar bears. Red pandas, once regarded as allied with bears, are not included because recent genetic analyses place them outside the Ursidae in a separate family more closely allied with the Procyonids, and show that their resemblance to pandas is the result of convergent evolution. The remaining 15 chapters explore interactions between bears and humans and issues relating to their conservation and management. Six of the eight contemporary bear species (giant pandas, sun bears, sloth bears, Asiatic black bears, brown bears, and polar bears) are found in Eurasia; three in North America (North American black bears, brown bears, and polar bears); and one (Andean bears) in South America. Recent genetic phylogenies suggest that bears diverged from the ancestors of the Ursidae between 23 and 35 million years ago, while the giant panda and its extinct relatives diverged from other bears between 12 and 20 million years ago. Andean bears diverged from the ancestors of the other species between 7 and 13 million years ago, while the remaining species (sun bears, sloth bears, North American and Asiatic black bears, and polar bears) appear to have diverged from each other within the last 5 million years. There are striking ecological differences between bear species that have important implications for their behavior, evolution, and conservation. Habitats range from tropical forests (in sun bears) to montane woodlands and grasslands

(Andean bears), northern forests, woodlands and grasslands (black and brown bears), and the marine margins and sea ice of the Arctic (polar bears). The average size of annual ranges runs from less than 10 km2 in some populations of sun bears to hundreds of square kilometers in brown bears and tens of thousands in polar bears. Diets are exclusively vegetarian in giant pandas (99% of whose diet is constituted by bamboo), consist of varying combinations of fruit, roots, foliage, insects, and vertebrates in sun bears, sloth bears, black bears, and brown bears, while polar bears are specialized hunters living principally on three seal species and also scavenging the carcasses of cetaceans and walruses. The smaller species are partly arboreal while the larger ones are principally terrestrial, and polar bears are accomplished swimmers. While black bears, brown bears, and Asiatic black bears (only in northern latitudes) hibernate in dens in winter (although the duration of denning varies with latitude), Andean bears, sun bears, sloth bears, giant pandas, and polar bears are active throughout the year. As in many other groups of mammals, the available information varies widely between species. Of all the bear species, North American black bears are the most widespread and most easily accessible to scientists, habituating relatively easily to observers, and long-term, individual-based studies of black bears provide detailed records of the lives of large numbers of recognizable individuals that make it possible to explore relationships between variation in ecology, development, and behavior. They demonstrate the benefits of long-term, individual-based studies and help to shed light on ecological processes that probably occur in many of the other species for which less complete information is available. Because of their larger ranges and lower densities, studies of brown bears and polar bears face greater logistical problems, but here, too, long-term studies are starting to provide extensive information on the lives of individuals and on the effects of differences in feeding behavior and hunting tactics on development, survival, and reproduction. In contrast, much less is known of the lives of Andean bears, sun bears, and sloth bears, whose habitat and activity patterns make detailed studies on individuals more difficult – and further research on their ecology and behavior is badly needed.

xix

Foreword

The lifestyles and behavior of all eight bears show substantial similarities. Adults usually forage alone or (in the case of females) with dependent young and occupy overlapping home ranges, which are typically larger for males than females. During breeding seasons, males range widely in search of receptive females. The prolonged dependence of juveniles on their mothers reduces the availability of receptive females, skewing the ratio of adults ready to breed towards males and generating intense competition between males for mating opportunities and strong selection for increased body size in males. Implantation is delayed, litter size is small (1–3 cubs, exceptionally 4), and the young are guarded and provisioned by mothers until they are ready to disperse. Male infanticide has been observed in several species and can represent a relatively common cause of cub mortality. Offspring adopt separate ranges from their mothers as they develop and, as in many other mammals, sons typically move further away than daughters, although the situation is reversed in giant pandas. Because of their fabled strength and ferocity and their upright stature when threatened, bears have had a prominent position in human religion, mythology, and iconography for over 2000 years. All species apart from giant pandas have been known to attack humans and they are widely feared, but attacks are uncommon, are very seldom predatory, and mostly occur when individuals are threatened or cornered or when mothers with dependent young are surprised. Experts recommend that the safest response for humans threatened by bears is to lie face down with hands behind your head and legs apart, so that you cannot be flipped over, and wait until the bear leaves. In contrast, cases of predation on domestic animals and exploration of human waste are much more common, especially in black and brown bears, stimulating hunting to control local problems. Bears are still hunted for sport in many countries, including North America, where black bears are a common target species in many US states. Throughout much of human history, they have also played a central role in popular entertainment. Since before Roman times, captive bears have been used in staged fights with dogs throughout much of Eurasia, and although bear baiting is now illegal in most countries, it still persists in a few. In addition, captive bears have commonly been taught to dance and have been displayed at fairs and games, and bear parts and products feature prominently in Chinese traditional medicine. Bears continue to be hunted and captured for both purposes.

xx

Conservation biologists often (and rightly) complain of the lack of connection between fundamental studies of the ecology and behavior of mammals and the needs of conservation biologists and managers for guidance on the optimal form of conservation or management strategies. They can have no such complaints about this book, for a large part of its second half is devoted to analyzing the direct and indirect effects of human impact on bear populations and assessing different ways of reducing human impact on existing bear populations. Chapters describe how overhunting, poaching, and reductions in their natural habitat have reduced the historical ranges of all bear species and how increasing human populations and changes in land use now threaten the persistence of many populations. In addition, human-induced changes in climate jeopardize the future of both pandas and polar bears. In polar bears, the impact of rising winter temperatures on the extent of sea ice affects their ability to hunt seals and has already led to substantial reductions in their numbers. Predictions suggest that, if current changes in winter temperatures continue, many polar bear populations are unlikely to persist to the end of this century. Climate change also threatens the remaining populations of giant pandas, as the distribution of over a third of the bamboo species that they rely on is likely to be affected by rising temperatures, and a substantial part of the protected areas set aside for pandas may no longer be suitable habitat by the end of the century. As a result, there is now an urgent need to identify and protect suitable habitat to replace them. At a time when the rate of publication is rising rapidly and scientists are relying more and more on electronic sources, large, taxonomically focused books might initially appear superfluous. However, the increasing volume of available information and the diversity of its origins and reliability make the need for expert syntheses of the available information more pressing. By bringing together authoritative, multi-authored reviews of the information available for different species, integrative books like Bears of the World make it possible to identify contrasts and similarities in the biology and ecology of different species and to compare the experience of different strategies for conserving and managing remaining populations, providing a clear view of the woods as well as the trees. Tim Clutton-Brock University of Cambridge

Acknowledgments

For their continuous effort, patience and dedication, as well as their beautiful drawings, we are extremely grateful to our illustrators Luciano Toma, Francesco Rinaldi, Enrico Chiarelli, Alberto Mattia Perronace, and Giulia Bombieri. We are indebted to Daniele Baisero for his great dedication and patience, and also for making the distribution maps of the eight bear species. Dr. Craig Hilton-Taylor, Head of the Red List Unit of IUCN, granted us permission to use the IUCN data for the distribution maps, a huge thanks to him! We would like to warmly thank the Cambridge staff for their support, in particular Dominic Lewis, Commissioning Editor, for his advice during the different stages of the book, Aleksandra Serocka, Senior Editorial Assistant, and Jenny van

der Meijden, Senior Content Manager, for their continuous help and useful advice throughout the entire preparation of this volume. Working with them has been as smooth as one could hope for, and their help has been invaluable throughout the entire process. We also thank the Press Syndicate at Cambridge University Press for approval of our title. We are also grateful to Yorgos Mertzanis and two anonymous reviewers who helped us to improve our proposal during the submission of the book plan to Cambridge University Press. Last, but not least, we are always in debt to our families for their continuous support.

xxi

Introduction Vincenzo Penteriani and Mario Melletti I consider the mountainside a special place, a place with power, as I do certain other valleys and basins . . . where grizzlies still roam. I return to these places year after year, to keep track of the bears and to log my life. The bears provided a calendar for me when I got back from Vietnam, . . . I had trouble with a world whose idea of vitality was anything other than the naked authenticity of living or dying. The world paled, as did all that my life had been before, and I found myself estranged from my own time. Wild places and grizzly bears solved this problem. Doug Peacock, 1990, Grizzly Years

Bears have fascinated people since ancient times. The relationship between bears and humans dates back tens of thousands of years, during which time we have also competed with bears for shelter and food. Our strong link with bears is also attested to by the Neanderthal burial of “Le Regourdou,” in France, where the skeleton of a Neanderthal in a fetal position was found under a funeral slab surrounded by the bones of a brown bear, probably sacrificed for the burial. Bears were also represented in rock paintings in caves inhabited by our ancestors in Europe. The bears depicted by our ancestors were cave bears, which roamed Eurasia until about 24,000 years ago when they became extinct during the Last Glacial Maximum. Recently, gene flow between extinct cave bears and brown bears has been discovered, providing direct evidence for ancestral hybridization between the two species which resulted in the modern Ursus arctos that we all know (Chapter 1). In human culture, bears also represent an important figure in Native American mythologies. For example, the bear is a symbol of power and strength. In fact, warriors of some tribes wore necklaces of bear claws. Bears also play a major role in several religious ceremonies in many North American tribes, which used to have a bear dance as part of their tribal traditions, and they represent an important clan animal in some native cultures, for example in tribes such as the Cherokees, Creeks, Hurons, and Navajos. Furthermore, bears are often found carved on totem poles of several tribes of north-western North America. Bears have also influenced the culture of many tribes in Asia. In fact, they are important animals for some tribes in Siberia, and the people of the Hokkaido and Ryukyu islands in Japan. For example, the Ainu people in Japan consider the bear as the “Spirit of the Mountains.” In Russia, at a Fat’yanovo cultural site dated to around 1500 BC, necklaces made with bear teeth were found, and other Neolithic findings have been discovered as far north as Lake Ilmen, in the Russian

oblast of Nóvgorod. Several bear claws with bronze mounting, dated between the ninth and eleventh centuries, were also discovered among a Finno-Ugrian group located along the River Tsna (a river in the Tambov and Ryazan oblasts of Russia), whereas another Finno-Ugrian group in the Urals venerated the bear as a symbol of heroism. The images of bears in popular culture have helped them to become an icon that most people know and love. The most famous example is the teddy bear, which has been one of the most popular stuffed animals since the early 1900s and continues to be a favorite of children. Developed almost simultaneously by toymakers Morris Michtom in the US and Margarethe Steiff in Germany, and then named after US President Theodore “Teddy” Roosevelt, the teddy bear is an iconic children’s toy celebrated in stories, songs, and films. More recently, Baloo from The Jungle Book, Winnie the Pooh, Yogi and Bubu, and Masha and the Bear tell us that the strong link between people and bears, which started more than 80,000 years ago, continues today. At the present time eight bear species are recognized, from the very popular polar bear, giant panda, brown bear, and American black bear to the lesser-known Andean bear, sun bear, sloth bear, and Asiatic black bear. In Chapter 1, the authors follow a different taxonomy for sun bear (Helarctos malayanus) and Asiatic black bear (Ursus thibetanus), ascribing them to the genus Melursus. However, we recognize that further genetic and morphometric studies are still required in order to fully understand the taxonomy of these two bear species. Therefore, and pending further investigation on Melursus taxonomy, we will still refer in this book to the genus Helarctos for sun bear and Ursus for Asiatic black bear, acknowledging that they might deserve a different genus in future. The conservation issue facing some of these species is a big conservation challenge today. Human activities, population encroachment, and poaching in bear landscapes continue to

1

Introduction

represent serious threats for some bear species or populations. As an example, in more recent times, human–bear conflicts have been exacerbated by the increasing number of people sharing the same landscapes with bears. Such coexistence has engendered an increase in conflicts, such as damage to livestock, crops, and apiaries, as well as the fear of bear attacks (Chapters 15–17, 20, 28). Moreover, in some regions people continue to keep bears in terrible captive conditions to extract bile and other body parts, mainly in Asia where the trade in these kinds of products still flourishes (Chapter 26). For instance, in China and Vietnam, thousands of bears are kept in tiny cages and bred for their bile. Furthermore, an emerging threat to bear conservation is related to the effect of global warning. This is the case, for example, for polar bears, which are now more endangered than even a few decades ago due to the dramatic reduction of sea ice. In fact, polar bears depend on the ice shelf for feeding, breeding, and movement, and they can only persist where the temporal and spatial availability of sea ice provides adequate access to their marine mammal prey (Chapter 21). The panda, another iconic bear species, is also threatened by climate change. The panda is an extremely specialized species that relies on bamboo for 99% of its diet, occurs in a very restricted range, and has one of the lowest reproduction rates among bears (Chapters 6 and 21). The challenge that bear conservation and management represents around the world makes this book extremely important and timely because it provides informative and complete accounts of everything you want to know on bears to a broad audience. However, even if our general knowledge of bear ecology and behavior has significantly increased in the past decades, we still have a lot to learn about this group of species, and in particular tropical bears (Chapters 4, 7–10, 22, 24–28). The idea to edit a major volume on the ecology and conservation of bears started in 2017 from the need to provide a comprehensive book on all the bear species that inhabit our planet, as well as a useful tool for both the general public and people more directly involved in the fields of animal ecology, behavior, and conservation such as researchers, wildlife managers, conservationists, stakeholders (such as farmers and hunters), and students. In our minds, this book would also been able to demonstrate why the study of human–bear interactions, and stakeholder perception and involvement, are crucial for bear conservation and management, with detailed examples and case studies from all the continents inhabited by bears. Such a book would also highlight the urgency of conservation actions that need to be put into practice in the different regions of the world, mainly for lesser-known bear species in developing countries. Additionally, and despite the long-term interest in bears, as well as the many groups

2

working on bears for decades and the large literature on these species around the world, a comprehensive and very detailed book on all the bear species in the world had never been published before. To make this idea a reality, we started by contacting more than 250 people among the best bear biologists in the world. Many of them replied enthusiastically to our invitation, saying that such a book was very much welcomed and long overdue. Of course, this positive feedback gave us further motivation to move forward with this project, in which 200 authors ultimately participated, many of them having spent their entire lives studying bears. The authors come from 33 countries spread across five continents, and work at very diverse institutions, such as research centers, universities, IUCN SSC Bear Specialist Group and Polar Bear Specialist Group and IBA, and non-governmental organizations (NGOs). To give you an idea of the heterogeneity of the book’s contributors, people involved in this project are from (in alphabetical order): Bhutan, Bolivia, Canada, China, Croatia, Denmark, Ecuador, Finland, France, Germany, Greece, Hungary, India, Iran, Italy, Japan, Malaysia, Mexico, Nepal, the Netherlands, Norway, Pakistan, Poland, Romania, Russia, Slovenia, Spain, Sri Lanka, Sweden, Turkey, UK, USA, and Venezuela. Together with the most important information on the ecology and behavior of bear species, the volume also includes specific chapters on taxonomy, phylogeny and genetics, population status and trends, as well as conservation status, management, and climate changes. The book is composed of 28 chapters subdivided into four sections: Part I – Systematics, Ecology, and Behavior (Chapters 1–5); Part II – Species Accounts (Chapters 6–14); Part III – Human–Bear Coexistence (Chapters 15–19); and Part IV – Conservation and Management (Chapters 20–28). We hope that you will enjoy this book at least as much as we have enjoyed its long preparation and our close collaboration with chapter contributors, and that the huge effort made by all the authors will be appreciated by the public and scientific community. We will consider that our work has achieved its goal if it is rewarded by an increase in the understanding of bears and their effective conservation. However, the future of bear species will depend on our capacity to find pragmatic solutions that should represent a trade-off between human growth and the needs of bears and their habitats. What bears, among the most charismatic creatures on the planet, do for people has an inestimable value for our most intimate life, and their loss will create a void impossible to fill. Being in a bear country captivates our minds and, at the same time, offers a lesson in humility by giving us the feeling that something more powerful than us is out there.

Part I Chapter

1

Systematics, Ecology, and Behavior

Systematics, Evolution, and Genetics of Bears Andrew C. Kitchener, Eva Bellemain, Xiang Ding, Alexander Kopatz, Verena E. Kutschera, Valentina Salomashkina, Manuel Ruiz-Garcı´a, Tabitha Graves, Yiling Hou, Lars Werdelin, and Axel Janke

Introduction Bears, Family Ursidae, are among the largest of the Carnivora, but the Ursidae is among the least speciose carnivoran family with only eight extant species. Owing to the bears’ widespread distribution in Eurasia, North America, and north-western South America, they occupy many habitats from the northern polar ice cap to tropical rainforests. Despite having few species, relationships between the Ursidae and other Carnivora and between ursid species (and indeed what constitutes an ursid) have been controversial. New molecular genetic techniques in recent years have allowed relationships to be explored in new ways, leading to some clarification, but debates continue owing to incomplete lineage sorting and ancient hybridization. Bears are threatened by humans; conflict with human land uses and exploitation for fur and other body parts, coupled with habitat loss and extirpation through direct hunting, have led to fragmentation of populations of all species. Molecular genetics are key to understanding current and historical relationships between isolated populations, including species’ colonization during glacial–interglacial cycles, to determine viability of local populations, needs for habitat corridors, and other aspects of population management, especially where bears are harvested for sport, etc. As natural habitats shrink, some bear species will inevitably require high levels of management, perhaps combining captive and wild populations following the IUCN’s One Plan Approach. In this chapter we review the systematics of the Ursidae and its relationships with other Carnivora, the molecular phylogenetics of extant ursid species, the phylogeography of and morphological variation within each species, and the use of molecular genetics to monitor bear populations for management and conservation.

Systematics of the Ursidae Wozencraft (1989) summarized the systematic history of the Carnivora, hence only a brief overview is given here. The classification of the Order Carnivora based on the morphology of the basicranium began with Turner (1848), was further developed by Flower (1869), and has since been refined, e.g. by Hunt (1974), who characterized the development and evolution of auditory bullae. The Carnivora is characterized by

morphological characters, including carnassial teeth, i.e. the fourth upper premolar (P4) and first lower molar (m1), which form two shearing blades for slicing through body tissues, and the presence of three elements in the auditory bulla: the caudal and rostral entotympanic and ectotympanic bones (Hunt 1998; Wozencraft 1989). Based on basicranial characteristics, the Carnivora comprises two suborders: the Feliformia (felids, viverrids, hyaenas, and related families) and the Caniformia (canids, ursids, pinnipeds, and musteloids). The Caniformia comprises two Infraorders, the Cynoidea (or Canoidea), including the Canidae, and the Arctoidea, including the remaining caniforms. Rose (2006) characterized the Arctoidea as having a suprameatal fossa (a hollow in the dorso-lateral wall of the middle-ear cavity) and loss of the third upper molar (M3), and that most arctoids have a single-chambered auditory bulla comprising mostly the ectotympanic bone. Some relationships within the Arctoidea remain controversial, although a consensus is growing. The main contentious areas include the relationship between the giant panda, Ailuropoda melanoleuca, the red panda, Ailurus fulgens, and the Ursidae (see below), whether the superfamily Pinnipedia is monophyletic, and the relationships between pinnipeds, the Ursidae, and Superfamily Musteloidea. Traditionally, the systematics of the Carnivora were based on morphological studies of extant and fossil taxa (e.g. Wozencraft 1989), but new molecular techniques have clarified many relationships, especially where homoplasy or convergence gave ambiguous or incorrect affinities between taxa. However, some relationships remain uncertain, owing to incomplete lineage sorting and ancient hybridization. The problematic systematics of the Arctoidea are discussed below. The Superfamily Pinnipedia includes seals (Family Phocidae), walrus (Family Odobenidae), and sea lions and fur seals (Family Otariidae). Morphological studies since the nineteenth century suggested a sister relationship between the Ursidae and Otariidae/Odobenidae based, for example, on the structure of the auditory bulla, and between the Mustelidae and Phocidae, based, for example, on the absence of an alisphenoid canal, thereby supporting a diphyletic origin of the Pinnipedia. However, recent morphological studies demonstrated a monophyletic Pinnipedia (e.g. Berta et al. 2015, but see Koretsky et al. 2016).

3

Systematics, Ecology, and Behavior Table 1.1 Classification of the extant Ursidae (Tedford 1976; Bryant 1996; Wagner 2010).

Class Mammalia Linnaeus, 1758; 1 Order Carnivora Bowdich, 1821; 33 Suborder Caniformia Kretzoi, 1943; 16, 194 Infraorder Arctoidea Flower, 1869; 15 Superfamily Ursoidea Batsch, 1788; 110 Family Ursidae Batsch, 1788; 110 Subfamily Ursinae Batsch, 1788; 110 Genus Ursus Linnaeus, 1758; 47 Ursus arctos Linnaeus, 1758; 47 Ursus maritimus Phipps, 1774; 185 Ursus americanus Pallas, 1780; 5 Genus Melursus Meyer, 1793; 155 Melursus ursinus (Shaw, 1790; pls. 58–59) Melursus malayanus (Raffles, 1821; 254) Melursus thibetanus (G. Cuvier, 1823; 325) Subfamily Arctotheriinae F. Ameghino, 1902; 236 Genus Tremarctos Gervais, 1855; 20 Tremarctos ornatus (F. Cuvier, 1825) Subfamily Ailuropodinae Grevé, 1894; 217 Genus Ailuropoda Milne-Edwards, 1870; 342 Ailuropoda melanoleuca (David, 1869; 12–13)

There is still no consensus on the relationship between the Ursidae and other Arctoidea. Several studies showed a sister relationship between the Ursidae and the Pinnipedia, but excluded the Musteloidea (i.e. Mustelidae, Procyonidae, Ailuridae, and Mephitidae) (Luan et al. 2013), while in most others the Ursidae was basal with the Pinnipedia and the Musteloidea grouped together (Eizirik et al. 2010; Nyakatura & BinindaEmonds 2012; Doronina et al. 2015) (Figure 1.1), or grouped the Ursidae with the Musteloidea (Luan et al. 2013). The Ursidae comprises eight extant species, which many authors consider to form two subfamilies (Wozencraft 1989; Goswami 2010; Wagner 2010): the Ailuropodinae includes the basal and highly derived giant panda, Ailuropoda melanoleuca, which is adapted to feeding on bamboo; and the Ursinae, which comprises two tribes, the Tremarctini, including the Andean bear, Tremarctos ornatus, and the Ursini, comprising the remaining six ursid species. Hunt (1998) divided the Ursidae into four subfamilies: the Ursinae containing the extant bears; the Ailuropodinae containing the giant panda; and two extinct subfamilies, the Amphicynodontinae and the Hemicyoninae. He further divided the Ursinae into three tribes – Ursini, Tremarctini, and the ancestral Ursavini. Most recent authors place the Andean bear in the subfamily Tremarctinae (e.g. Soibelzon et al. 2005; Mitchell et al. 2016), but Wagner

4

(2010) suggested that the Tremarctinae Merriam and Stock, 1925 is pre-dated by the Arctotheriinae Ameghino, 1903, but it actually dates to Ameghino (1902) (Table 1.1). Currently, the extant Ursinae includes three genera: Helarctos, for the sun bear, H. malayanus; Melursus, for the sloth bear, M. ursinus; and Ursus, comprising the brown bear, U. arctos, polar bear, U. maritimus, Asian black bear, U. thibetanus, and American black bear, U. americanus. Previously, three genera were recognized for Asian bears, including Melursus (sloth bear), Selenarctos (Asian black bear), and Helarctos (sun bear). A recent molecular phylogenetic study by Kumar et al. (2017) showed that the sloth and sun bears are sister species in a clade with the Asian black bear, which is basal. This suggests that either all Ursinae should be in the genus Ursus as proposed previously (e.g. Wozencraft 1989; Hunt 1998) or that the three Asian bears may be in a separate genus, Melursus. Recently, Kitchener et al. (2017) employed Hennig’s Rule, which discriminates genera if divergence times are c.5 million years or more. Using Hennig’s Rule divides the Ursinae into Melursus for Asian bears and Ursus for the rest (Table 1.2). The Ursidae has a primitive dental formula (I3/3 C1/1 PM4/4 M2/3) and a Type A auditory bulla (Hunt 1974), comprising mostly the ectotympanic but also rostral and two

Systematics, Evolution, and Genetics of Bears Table 1.2 Classification of the Ursidae, including extinct taxa.

Family Ursidae Batsch, 1788 Subfamily Hemicyoninae Frick, 1926 Tribe Cephalogalini Bonis, 2013 Adelpharctos Bonis, 1971 Cyonarctos Bonis, 2013 Phoberogale Ginsburg & Morales, 1995 Filholictis Bonis, 2013 Cephalogale Jourdan, 1862 Tribe Phoberocyonini Ginsburg & Morales, 1995 Plithocyon Ginsburg, 1955 Phoberocyon Ginsburg, 1955 Tribe Hemicyonini Frick, 1926 Zaragocyon Ginsburg & Morales, 1995 Dinocyon Jourdan, 1861 Hemicyon Lartet, 1851 Subfamily Ursavinae Hendey, 1980 Ballusia Ginsburg & Morales, 1998 Ursavus Schlosser, 1899 Subfamily Agriotheriinae Kretzoi, 1929 Agriotherium Wagner, 1837 Subfamily Ailuropodinae Grevé, 1894 Tribe Indarctini Abella et al., 2012

Figure 1.1 Phylogenetic tree showing a version of the relationships among the Arctoidea (Koepfli et al. 2017).

Miomaci Bonis et al., 2017 Indarctos Pilgrim, 1913 Tribe Ailuropodini Grevé, 1894 Kretzoiarctos Abella et al., 2012 Agriarctos Kretzoi, 1942 Ailurarctos Qi et al., 1989 Ailuropoda Milne-Edwards, 1870 Subfamily Arctotheriinae Ameghino, 1903 Plionarctos Frick, 1926 Arctodus Leidy, 1854 Arctotherium Burmeister, 1879 Tremarctos Gervais, 1855

Figure 1.2 Dissected auditory region of a bear, showing the internal carotid artery looped within the inferior petrosal venous sinus on the lateral edge of the basioccipital bone. AL, alisphenoid; BO, basioccipital; BS, basisphenoid; PE, petrosal (Hunt 1977).

Subfamily Ursinae Batsch, 1788 Ursus Linnaeus, 1758 Melursus Meyer, 1793

caudal entotympanic bones (Hunt 1998). The internal carotid artery loops inside the petrosal venous sinus, covered by the basioccipital, and may act as a countercurrent heat exchanger to cool the brain (Hunt 1998) (Figure 1.2).

The Ursinae and Arctotheriinae share several morphological characters; a reduced lacrimal forming a vestigial bony rim around the naso-lacrimal foramen; a flat auditory bulla, which is much smaller than the mastoid, squamosal, and basioccipital processes; reduced premolars, whereby P4 has lost most of its shearing function; the upper and lower molars are quadrate (with four main cusps: paracone, metacone, protocone, and a large metaconule, which migrated to the

5

Systematics, Ecology, and Behavior

postero-internal border to form the molars’ oblong shape), and the rear border of M2 is greatly enlarged and elongated, forming a distinctive heel and making it the longest tooth in the maxilla (Wozencraft 1989; Hunt 1998). The enamel on the surface of the talonid and trigonid basins of m2 of arctotheriines is tuberculated and forms a “z” pattern on the talonid (Hunt 1998). The skull is shorter than those of the Ursinae and there is a pre-masseteric fossa in the mandible in Tremarctos and Arctodus, but not Plionarctos. The diploid (2n) chromosome number is 52 biarmed autosomes in Tremarctos compared with 2n = 74 mostly acrocentric chromosomes of the Ursinae, and 2n = 42 mostly biarmed autosomes in the giant panda (Chorn & Hoffmann 1978). The giant panda’s classification was controversial since its description by David (1869), who regarded it as a bear, but Milne-Edwards (1870) classified it as a procyonid. Later authors proposed placing it in a family (Ailuridae) with the red panda, or in a monotypic family (Ailuropodidae). The controversy continued as different morphological features were presented for and against the two main hypotheses until molecular studies demonstrated unequivocally that the giant panda is basal to the Ursinae and Arctotheriinae. It is usually classified in the Ailuropodinae, although some still place it in its own family (although not recently, e.g. Thenius 1989), owing to its highly derived morphology and deep divergence time of c.12.5 million years ago (Ma) (Kutschera et al. 2014). A key morphological character of the Ailuropodinae is a greatly enlarged radial sesamoid in the wrist, the so-called panda’s thumb, which grasps stems, such as those of bamboo. In most other extant ursids the radial sesamoid is not enlarged, but recently the Andean bear’s moderately enlarged radial sesamoid was described (Salesa et al. 2006). Therefore, enlarged radial sesamoids are either plesiomorphic (i.e. an ancestral character shared by two or more taxa), or they evolved independently in primarily herbivorous ursids. Other morphological characteristics of ailuropodines include the highly domed skull for attachment of massive jaw muscles, and enlarged molars and premolars for crushing tough bamboo stems and shoots.

Fossil History of the Ursidae There is no agreement regarding quite what a bear is, because opinion differs as to which basal taxa should be included in the Ursidae. For any mammal group the deeper one descends into the fossil record, the more similar taxa become, until the origin of the group is reached. Today’s bears differ greatly from those of the late Eocene and Oligocene (c.35–23 million years ago). Consequently, opinion differs as to which basal taxa should be included in the Ursidae (Table 1.2). A group of genera, including Amphicynodon, Amphicticeps, Parictis, Kolponomos, Allocyon, and Pachycynodon, distributed in Europe, Asia, and North America, constitute the amphicynodonts. This is either a paraphyletic stem lineage within the Ursidae or a monophyletic family Amphicynodontidae that may or may not be sister

6

taxon to the Ursidae. The latter view point has gained credence in recent years, but without consensus. Here, we consider the amphicynodonts to constitute a distinct family and exclude them from further discussion. The earliest Ursidae are from the earliest Oligocene. The family is subdivided into a number of subfamilies: Hemicyoninae, Ursavinae, Agriotheriinae, Ailuropodinae, Arctotheriinae, and Ursinae. The Hemicyoninae is the basal subfamily and encompasses about 10 genera in three tribes. Among these is the genus Cephalogale, often called “the earliest bear.” This genus has recently been revised (de Bonis 2013), which showed that traditional “Cephalogale” is an assemblage of variably related, primitive Ursidae, within the tribe Cephalogalini, which now comprises five genera – Cephalogale, Filholictis, Cyonarctos, Adelpharctos, and Phoberogale – of jackal- to wolf-sized, rather canid-like animals. Cephalogalini is known from the earliest Oligocene (c.34 Ma) to the early Miocene (c.19 Ma). Their geographical range extended from Western Europe to the Midwestern United States, with Cephalogale and Phoberogale occurring on both continents. De Bonis (2013, p. 810) notes that no antecedents of Cephalogalini are known from the Eocene and surmises that they (and the Ursidae) may have had an Asian origin, where Cephalogalini are poorly known. A second tribe of Hemicyoninae is the Phoberocyonini, revised by Ginsburg and Morales (1998). This tribe comprises Phoberocyon (not to be confused with Phoberogale) and Plithocyon, two genera with wide geographical ranges, both known from France to Florida. Phoberocyon is the older genus, with a stratigraphic range encompassing the late Oligocene to late middle Miocene (c.29–14 Ma), whereas Plithocyon is restricted to the middle Miocene (c.16–11.5 Ma). These two genera include larger, more derived species than the Cephalogalini, with the largest, e.g. Phoberocyon aurelianensis, within the size range of ursids today. Although phylogenetically close, Phoberocyon evolved toward hypercarnivory, with taller cusps on the lower carnassial and a narrow, trenchant talonid, whereas Plithocyon became more hypocarnivorous, with a low, wide carnassial. The reasons for this ecological divergence are debatable and complicated by the presence of both genera on multiple continents, each with different competitors. Nevertheless, the Ursidae would never be as hypercarnivorous again. The Hemicyonini comprises two genera, Hemicyon and Dinocyon. Hemicyon includes generalized, medium- to largesized hemicyonines of early and middle Miocene age with a broad geographical range from western Europe (where it is relatively common) to the midwestern United States, although Hunt (1998) suggests that the latter (Hemicyon barbouri) represents a distinct genus. Hemicyon is also recorded from Kenya by a single upper carnassial from the early Miocene (SchmidtKittler 1987). This tooth is the only record of a bear from Africa prior to Agriotherium in the latest Miocene (see below), despite decades of searching at productive localities, raising many presently unanswerable questions. Dinocyon is much less common than Hemicyon. It is exclusively known from western Europe, in sediments dating from

Systematics, Evolution, and Genetics of Bears

the middle Miocene to the earliest late Miocene (c.16–11 Ma). Dinocyon includes some of the largest bears of all time and, at least in tooth dimensions, rivaled giants such as Agriotherium africanum and Arctotherium angustidens (see below). Closer to crown-group Ursidae, relationships between subfamilies are a little better understood (Abella et al. 2012), although the first, Ursavinae, is clearly a paraphyletic stem lineage. It comprises the genera Ballusia and Ursavus. The former is exclusively European from the early Miocene, whereas the latter comprises several species from the early to late Miocene of Eurasia and the middle Miocene of North America. Ursavus is traditionally considered ancestral to more derived bears. They are all small to medium-sized animals with generalized omnivore characteristics. Following the Ursavinae is the subfamily Agriotheriinae, which includes only the genus Agriotherium. A. africanum is the first unequivocal bear to be described from sub-Saharan Africa, by Hendey (1980). Until that time, bears were only thought to have reached North Africa in the late Pleistocene, but never to have crossed the Sahara. Agriotherium africanum was considerably larger and more carnivorous than modern Ursus. Crown-group Ursidae (the common ancestor of all living bears and all of its descendants) includes three subfamilies. The Arctotheriinae encompasses the South American ursid radiation, with Tremarctos ornatus as its surviving representative. The arctotheriines have antecedents in North America, where the ancestral taxon is Plionarctos of the late Miocene to Pliocene. Tremarctos also originated in North America, where T. floridanus overlapped in time with Plionarctos. However, most characteristic of the arctotheriine radiation are the short-faced bears, Arctodus (Pliocene of North America) and Arctotherium (Pleistocene of South America). Among the latter, Arctotherium angustidens may have been the largest bear (and carnivoran) of all time, with an estimated body mass of up to 1500 kg (Soibelzon & Schubert 2011). Curiously, the South American short-faced bear was more closely related to the living Andean bear (which is not short-faced) than it was to the North American short-faced bear. This is a striking example of convergence in size and morphology to ecological circumstance. The Ursinae includes the extant genera Melursus and Ursus. The antecedents of Melursus are poorly known except for the Asian black bear. However, the genus Ursus is extremely well known in the fossil record, particularly the cave bear, U. spelaeus, which is studied from many thousands of specimens across large parts of western Eurasia, where it hibernated (and died) in caves. Numerous other species of Ursus have been described, including fossil relatives of the American black bear. The final subfamily, the Ailuropodinae, includes the giant panda and relatives. The palaeontology of this subfamily has progressed greatly recently (Abella et al. 2012; de Bonis et al. 2017). The giant panda, once thought to be very isolated phylogenetically among the Ursidae, is now known to be among a large diversity of genera and species in the Ailuropodinae. Two tribes are recognized, the Indarctini and the

Ailuropodini. The former includes the early genus Miomaci of the late Miocene in Europe and Indarctos, a genus known from the late Miocene of Eurasia and North America. The Ailuropodini comprises several genera, the oldest and most primitive of which are the European Kretzoiarctos and Agriarctos of late middle and early late Miocene age. Together with the Chinese Ailurarctos, these genera form a clade that is sister to Ailuropoda, culminating in the extant giant panda. This lineage shows a strong trend toward increased size and specialization to herbivory.

Molecular Phylogenies The relationships among today’s bear species have been unclear or contradictory depending on which methods were used. In particular, the relationships between the six ursine species remained unresolved until recently, because early molecular studies contradicted each other, mostly because of insufficient data (Talbot & Shields 1996; Yu et al. 2004). However, even larger molecular data sets that included nuclear genes did not confidently resolve the bears’ phylogeny (Pages et al. 2008; Kutschera et al. 2014), resulting in a “forest of gene trees” with contradicting phylogenies. This suggested that bear evolution was either shaped by incomplete lineage sorting or that ancient hybridizations complicated phylogenetic reconstruction (Kutschera et al. 2014). The first analyses of whole-genome data from all ursine and arctotheriine species have recently resolved relationships among bears (Figure 1.3; Kumar et al. 2017). Previously, the position of the American and Asian black bears with respect to their relatives was much debated (Pages et al. 2008; Kutschera et al. 2014). In the bifurcating tree (Figure 1.3), the American black bear is the sister species to polar and brown bears, while

Figure 1.3 Phylogenomic analyses of 18,621 maximum likelihood trees estimated from non-overlapping 100-kb genome fragments from whole genome data of seven bear species (Kumar et al. 2017). (A) Schematic of coalescent species tree with 100% bootstrap support of all branches. Divergence time estimates in million years ago (Ma) from Kumar et al. (2017) were mapped onto the tree that was obtained from an analysis of 5.2 million bp coding sequences in MCMC tree (PAML) with multiple calibration points. (B) Split network analysis at a 7% threshold level. Ma, million years ago; ABC, brown bear from the Admiralty, Baranof, and Chichagof (ABC) Islands. Paintings by Jon Baldur Hlidberg (www.fauna.is). Modified from Kumar et al. (2017), figure 1.3, licensed under CC BY 4.0.

7

Systematics, Ecology, and Behavior

sun and sloth bears are grouped with the Asian black bear. The Arctotheriinae forms the sister group to all ursine bears. The giant panda was consistently placed as sister group to the Ursinae/Arctotheriinae in previous phylogenetic analyses (e.g. Kutschera et al. 2014). However, analyses of bear genomes have shown that gene flow among species typifies the bears’ evolutionary history (Kumar et al. 2017). Although harder to interpret, the bears’ evolutionary radiation is actually a network, where signals from deviating phylogenetic trees become evident, with the American black bear placed between Asian bears and polar/ brown bears (Figure 1.3B; Kumar et al. 2017). A brown bear from the Admiralty, Baranof, and Chichagof (ABC) Islands, off the western Canadian coast, is positioned between polar and brown bears, consistent with gene flow between ABC brown bears and polar bears (Miller et al. 2012; Cahill et al. 2015). This network represents best the genome-wide evolutionary signal among extant ursids (Bapteste et al. 2013), because frequent hybridization among bears means their genomes are mosaics of evolutionary histories. A similar mosaic can result from incomplete lineage sorting, where ancestral gene lineages are randomly sorted along the species tree, and where phylogenetic signals may also differ from the species tree (Pamilo & Nei 1988). Only if all phylogenetic information in the evolutionary network is forced into one bifurcating tree is the more conventional view of bear evolution recovered (Figure 1.3A). While this tree may serve many purposes, such as taxonomy, conservation genetics, or molecular clock analyses, it provides an incomplete understanding of the bears’ evolutionary history and it may also explain why some morphological characters are difficult to place on a simple, bifurcating tree. Genetic differences among the bears can be converted into time, assuming that substitutions accumulate relatively constantly between species, as in a molecular clock (Zuckerkandl & Pauling 1962). Molecular divergence times among bears have been studied extensively (e.g. Kutschera et al. 2014; Kumar et al. 2017), with particular interest in the divergence of polar and brown bears (e.g. Hailer et al. 2012; Miller et al. 2012). Genome-wide estimates suggest the deepest split in the Ursinae (Ursus versus Melursus) was 5.0 (4.5–6.0) million years ago (Ma), and the divergence of polar and brown bears was 0.9 (0.6–1.1) Ma (Kumar et al. 2017). Phylogenetic reconstruction of bears exemplifies the complications that gene flow creates for understanding evolutionary relationships. One of the first molecular studies on bear species used mitochondrial DNA (mtDNA) to study the relationship between polar and brown bears (Cronin et al. 1991). This analysis suggested that polar bears originated from brown bears from the ABC Islands, questioning their validity as a distinct species and leaving Ursus arctos paraphyletic. However, some 20 years later, a comprehensive analysis of nuclear gene loci demonstrated that polar bears are a separate, ancient lineage that diverged from the common ancestor of polar and

8

brown bears about 600,000 years ago (ya) (Hailer et al. 2012). Therefore, the mtDNA of ABC Islands brown bears probably transferred to polar bears by hybridization, making it the first example of gene flow among bear species. For this transfer of mtDNA, female brown bears and male polar bears must have hybridized some 150,000 ya, coinciding with the penultimate glacial period (c.185,000–135,000 ya; Marine Isotope Stage 6; Margari et al. 2010). The brown bear mitochondrial haplotype has been identified in hundreds of polar bears studied for this locus so far (e.g. Cronin et al. 1991; Hailer et al. 2012). Under this scenario, the brown bear mitochondrial haplotype spread to all polar bears in a comparatively short time, also implying that brown bear mtDNA was selectively advantageous for polar bear survival, or the population bottleneck was very severe. An alternative explanation for high genetic similarity of polar and brown bear mtDNA is that gene flow occurred from polar bears into some brown bear populations (Cahill et al. 2013, 2018; Liu et al. 2014; Hassanin 2015). Indeed, gene flow between polar and brown bears was repeatedly detected in analyses of entire genomes (e.g. Miller et al. 2012; Cahill et al. 2015; Kumar et al. 2017). Up to 8.8% of the brown bear genome is of polar bear ancestry, but polar bear genomes harbor no significant genetic introgression from brown bears (Cahill et al. 2015). Cahill et al. (2018) suggested that brown bears carried some polar bear alleles away from where admixture occurred through male-dominated dispersal. It is unclear if this gene-flow asymmetry reflects continuing hybridization between the two species, is a remnant from the speciation event, or results from strong selection pressure on the highly specialized polar bear. Genome data from all bear species show that gene flow was much more common than assumed for mammalian genera (Kumar et al. 2017). Whether hybridization is current or ancestral among bears, different parts of their genomes have different evolutionary histories. This complicates our understanding of bear evolution. For example, some parts of the American black bear’s genome support a close relationship with the polar and brown bears, while other parts place the American black bear closest to the Asian black bear (Figure 1.3B). Phylogenetic reconstruction among the Asian bears is even more complicated and consistent with natural hybridization between Asian black bears and sun bears (Galbreath et al. 2008). Technical advances in analyses of ancient DNA allow studies of entire genomes of extinct species. Recently, a study found gene flow from extinct cave bears, Ursus spelaeus, into brown bears (Barlow et al. 2018), providing direct evidence for ancestral hybridization. Such interspecific gene flow might have been advantageous for adapting to changing environments.

Phylogeography and Intraspecific Taxonomy Although relationships between species are now better understood, relationships between populations within species and the taxonomic status of those populations are often unresolved in

Systematics, Evolution, and Genetics of Bears

the Ursidae despite descriptions of many different subspecies and even species. For example, 232 recent and 39 fossil species and subspecies of today’s Ursus arctos have been described (see Erdbrink 1953; Hall 1981), which Kurtén and Anderson (1980) considered “a waste of taxonomic effort, which, as far as we know, is unparalleled.” Phylogeographical studies allow better understanding of how populations are related to each other, where refugia existed during glaciations, and how species have recolonized continental ranges. However, lack of genetic and morphological data and their frequent apparent incongruence, along with high intraspecific variability, have led to poor resolution of intraspecific taxonomy and relationships in some species. For example, polar and Andean bears are considered monotypic, despite geographical variation in size and some genetic structuring of populations, although ESUs (evolutionary significant units) or MUs (management units) are recognized owing to genetically discrete populations, which occupy different habitats or experience different climatic conditions. In contrast, 8–12 extant brown bear subspecies are described; some are historically and widely recognized, but recent ones are not fully accepted, and for some, their geographical ranges are poorly known. Below, we briefly review the phylogeography and current status of subspecies within each bear species.

Ailuropoda melanoleuca The giant panda was widely regarded as monotypic until Wan et al. (2003) described A. m. qinlingensis, from the Qinling Mountains, Shaanxi Province, based on genetic differentiation from A. m. melanoleuca in Sichuan. This proposed new subspecies appears to have a browner pelage and smaller skull with larger molars (Wan et al. 2005). However, the distribution gap between the two putative subspecies is caused by human impacts, so genetic differences may represent genetic drift in now separate populations or clinal variation in a previously continuous range. The giant panda’s phylogeography shows a mismatch between mitochondrial and nuclear genetic studies. Early studies of short mtDNA sequences showed no haplogroups, even using longer sequence lengths up to 680 base pairs (bp) (Zhang et al. 2007; Hu et al. 2010a), which distinguish brown bear subclades. Recent studies, using complete mitochondrial genomes, revealed at least three extant and two extinct clades in giant pandas (Barlow et al. 2019). However, although mitochondrial clades are well defined and calibrated, they are not restricted to any population and show no geographical subdivision. In contrast, nuclear DNA reveals quite finely resolved geographical patterns with the Qinling Mountains, home to the most divergent population (Zhao et al. 2013), and Xiaoxiangling as a probable refugium (Chen et al. 2013). This discordance in genetic markers probably results from femalebiased dispersal (Hu et al. 2010b). Also, considering drastic range decreases and population fragmentation, it is probable that many other genetic lineages are extinct, which could have highlighted the species’ history and phylogeography.

Tremarctos ornatus Although subspecies were proposed previously (García-Rangel 2012), today the Andean bear is regarded as monotypic. However, genetic studies show that Andean bears comprise two ESUs: the Northern Andean clade (NAC; Venezuela, Colombia, Ecuador, and north-central Peru) and the Southern Andean clade (SAC; southern Peru and the northern and central Bolivian Andes), which split around 500,000 ya (Ruiz-García 2013; Ruiz-García et al. 2020a). Additionally, in Bolivia, some Andean bears from Santa Cruz Department were more related to the NAC than the SAC. These results and the slightly higher-level genetic diversity in the SAC show that the latter was the ancestral Andean bear population, contradicting palaeontological finds, which support the most northern Andean bear population as ancestral. Haplotypes vary in how they spread through the Colombian Andean Cordilleras. Gene-flow estimates were relatively high; therefore, geographical barriers have not prevented dispersal of Andean bears there. However, significant genetic heterogeneity was found in northern Colombian populations, with significant spatial autocorrelation for the Andean bear in Colombia (Ruiz-García et al. 2020b). In contrast, no genetic differentiation was found between different Cordillera Provinces, and northern– southern regions in Ecuador (Ruiz-García et al. 2020c). The first molecular genetic studies of Andean bears used nuclear markers (microsatellites; Ruiz-García 2003, 2007; Ruiz-García et al. 2005) and estimated low to medium genetic diversities. However, mitochondrial markers yielded high levels of genetic diversity. This paradox is probably explained by “ascertainment bias” (Ellegren et al. 1995), whereby the species for which microsatellites were developed shows higher genetic diversity than that of the non-target species. In this case, the microsatellites used for Andean bears were originally designed for American black bears (Kumar et al. 2017).

Melursus malayanus Usually two subspecies of sun bears are recognized: M. m. malayanus on the Asian mainland and Sumatra, and M. m. euryspilus from Borneo (Corbet & Hill 1992). Meijaard (2004) analyzed skull morphometrics of sun bears and confirmed that the Bornean population, which has a smaller skull and relatively longer upper tooth row, is distinct from Sumatran and mainland populations. Very little is known about sun bear genetics. Meijaard (2004) cited an unpublished mtDNA study by L. Waits on sun bears from Sumatra and Borneo that yielded five clades with no geographical structure, suggesting some gene flow between islands during the Last Glacial Maximum (LGM). From published data, two mtDNA lineages are reported from Borneo (divergence 5.6 ± 1.9 base pairs (bp)), but no geographical separation (Onuma et al. 2006). Yu et al. (2007) and Krause et al. (2008) obtained mtDNA sequences from three putative continental Asian sun bears that could represent another subspecies, and which yielded two additional

9

Systematics, Ecology, and Behavior

haplotypes that diverged from the two Bornean lineages by 14.39 ± 3.39 bp and 16.26 ± 3.57 bp, respectively. The differences between these mtDNA clades suggest a robust structure exists, dividing continental and island bears, but this is not yet fully resolved.

Melursus ursinus Usually two subspecies of sloth bear are recognized: M. u. ursinus from peninsular India and M. u. inornatus, which has a smaller skull, from Sri Lanka (Corbet & Hill 1992). However, sample sizes for cranial lengths were very small. There may be clinal variation in skull size with latitude from larger northern animals to smaller southern animals, but this has not yet been studied. The sloth bear’s phylogeography has not been studied. However, in GenBank there are seven sequences of mtDNA fragments, including Zhang and Ryder 1993 (captive); Yu et al. 2007 (no locality); Talbot and Shields 1996 (captive); Mohan et al. unpublished (India); and Krause et al. 2008 (no locality). Not all these sequences could be aligned, although a comparison reveals at least three highly divergent haplogroups, lacking any geographical information.

Melursus thibetanus Asian black bears are widely distributed in southern and eastern Asia; several subspecies are usually recognized, including M. t. thibetanus from Nepal to SE Asia, M. t. japonicus from Japan, M. t. formosanus from Taiwan, M. t. gedrosianus from Pakistan, M. t. ussuricus from the Russian Far East, Korea and northern China, M. t. mupinensis from southern and central China, and M. t. laniger from Kashmir and Punjab (Ellerman & Morrison-Scott 1953). The Asian black bear’s phylogeographical structure, based on mtDNA, is robust, but not yet fully resolved. It includes several haplogroups roughly congruent with seven putative subspecies (Hwang et al. 2008; Kadariya et al. 2018). The Japanese subspecies forms a distinct basal mitochondrial clade (Wu et al. 2015; Kadariya et al. 2018), which probably colonized Japan in the Middle Pleistocene, and is further substructured among the Honshu bears (Ohnishi et al. 2009; Kim et al. 2011). The mainland Asian clade probably represents a latitudinal cline, with increasing adaptation to more tropical climates from north to south, including the Taiwanese population, but the exact phylogenetic order of the matrilines is uncertain (Ohnishi et al. 2009; Kim et al. 2011; Wu et al. 2015; Kadariya et al. 2018). Recently, Kadariya et al. (2018) showed that the Himalayan population is a distinct basal lineage to mainland Asian populations. This suggests that three subspecies are recognizable: M. t. thibetanus from mainland Asia, M. t. laniger from the Himalayas, and M. t. japonicus from Japan. This species’ phylogeographical structure is consistent with female philopatry and it seems very likely that it was influenced by the Pleistocene and Holocene glaciations throughout its range.

10

Although the Ussuri black bear (known as M. t. ussuricus) is separated by more than 1500 km from other populations, it is not genetically distinct. The Ussuri population arose very recently, separating c.20,000 years ago (peak LGM), and adapting to colder climates. The distribution gap, cf. the tiger, Panthera tigris, arose even later, owing to forest loss since the early Neolithic, c.9000 years ago, and the exploitation of bears for traditional medicines by early civilizations in northern China since at least the Shang dynasty c.3500 years ago (Barnes 1999; Ren 2000). Owing to its ecological distinctiveness and geographical isolation, the Ussuri population should be considered as a separate MU.

Ursus americanus Owing to its widespread distribution, coupled with polychromatism (Rounds 1987), many subspecies of U. americanus (16 listed by Hall 1981) were recognized, with U. a. americanus occupying most of the species range. For American black bears phylogeographical structures are described for both mitochondrial and nuclear DNA (Wooding & Ward 1997; Pelletier et al 2011; Puckett et al. 2015). Analysis of mtDNA revealed two lineages; the continental lineage is subdivided into eastern and western subclades, while the coastal lineage has a much simpler structure along the Pacific coast of North America. Nuclear SNP (single nucleotide polymorphism) data also showed quite clear geographical structure, including three main continental nuclear clusters (Alaska, Eastern, and Western), which were divided into nine subdivisions (Puckett et al. 2015). However, nuclear and mitochondrial clusters mostly did not coincide. The current phylogeographical structure is believed to have originated before the LGM [31–67 kya (thousand years ago) for nuclear clusters and 169 kya to 1.07 Ma for the mitochondrial ones] in several refugia in northern and southern North America, including Beringia, the north-west Pacific, the Southwest, and the Southeast (Puckett et al. 2015). Neither mitochondrial nor nuclear structures match subspecies delineation very well (Puckett et al. 2015). Indeed, it is difficult to discern subspecies owing to complex admixture following expansion from these refugia, although Puckett et al. (2015) suggested that the nine subclusters could be associated with subspecies, which still seems too many for a highly mobile species showing extensive admixture. Perhaps the three mtDNA clusters, which appear mostly geographically discrete, would be better recognized as subspecies, including U. a. americanus (clade A-east) in the east, U. a. emmonsii in the west (clade B) and U. a. cinnamomum (clade A-west) from the central range.

Ursus maritimus The polar bear is a monotypic species, although subspecies were previously proposed. Wilson (1976) found clinal variation in skull morphometrics across North America, with the possible exception of a distinct population in South Alaska,

Systematics, Evolution, and Genetics of Bears

which was proposed as a subspecies. Currently, 19 MUs are recognized (Durner et al. 2018). For the polar bear, mtDNA reveals no clear genetic structure, although particular locations may differentiate in haplotype frequencies (Campagna et al. 2013; Kutschera et al. 2016). This lack of structure is partially attributed to long-distance movements of individuals and partially to the relatively recent origin/divergence of the species’ mtDNA lineages (see above). In contrast, microsatellite and SNP data show four to six geographical clusters, which are consistent between studies (Kutschera et al. 2016; Malenfant et al. 2016) and show differentiation between adjacent populations (Crompton et al. 2008; Campagna et al. 2013; Viengkone 2016). Putative populations are often identified as MUs, although microsatellite clusters have broader geographical ranges and may include several putative MUs. A comprehensive reconstruction of the species’ history, cf. brown bear and American black bear, is prevented by current unavailability of sufficient genetic markers with enough phylogenetic signal.

Ursus arctos The brown bear is the most widespread ursid species and, given the wide variety of habitats it occupies, it is unsurprising that its morphology is so variable. As a result, many subspecies have been recognized, but the principal extant ones are: Eurasian brown bear, U. a. arctos; Syrian bear, U. a. syriacus; Isabelline bear, U. a. isabellinus; blue bear, U. a. pruinosus; Kamchatkan bear, U. a. beringianus/piscator; Siberian brown bear, U. a. collaris; Japanese brown bear, U. a. lasiotus; Kodiak bear, U. a. middendorffii; and grizzly bear, U. a. horribilis (Ellerman & Morrison-Scott 1953). However, brown bears show much individual, seasonal, age, and sexual variation, and thus it can be challenging to ascertain subspecies taxonomy based on morphology alone. Although there have been many phylogeographical studies on brown bears, very few used nuclear DNA, so phylogeographical patterns often reflect female natal philopatry. The phylogeographical structure of the brown bear is well resolved. Several mitochondrial haplogroups are recognized, some of which are widely distributed and sympatric, whereas others are restricted to particular geographical regions. The picture is further complicated (see above), because polar bears share a mitochondrial DNA haplogroup with ABC Islands brown bears and an extinct Irish population (Edwards et al. 2011). Some studies found that current phylogeographical patterns do not match distributions of mtDNA haplotypes from fossil specimens (e.g. Paetkau et al. 1998; Barnes et al. 2002; Hofreiter et al. 2004; Valdiosera et al. 2007; Ersmark et al. 2019), such that today’s patterns result from female philopatry, population expansion and fragmentation, and genetic drift. MtDNA reveals at least six major clades grouped into two main branches. The first branch includes Western European (clade 1) and the Iranian clade from south-west Asia and the ABC Islands (clade 2; including the polar bear) (Davison et al. 2011;

Ashrafzadeh et al. 2016; Salomashkina et al. 2017). The second branch includes bears from southern North America and Hokkaido (clades 4 and 3d), Tibet (clade 5), and superclade 3, which occurs throughout most of the Eurasian and Northern American species range and is divided into two (3a and 3b) to four (3a, 3b, and two Caucasian) subclades (Murtskhvaladze et al. 2010; Çilingir et al. 2016; Salomashkina et al. 2017) (Figure 1.4). There is also a recently discovered Himalayan clade that appears to form a third main branch (Lan et al. 2017). Particular traits of the brown bear’s phylogeography allow reconstruction of its complex colonization history, which was heavily influenced by glacial cycles, involving multiple cycles of range contractions followed by lineage sorting and subsequent range expansions (Davison et al. 2011; GarcíaVázquez et al. 2017; Anijalg et al. 2018). Therefore, care should be taken in making taxonomic decisions based on contemporary genetic data alone without a deeper time perspective. MtDNA reflects matrilinear history, but it is not directly connected with environmental adaptation, and thus in some cases animals inhabiting the same area may morphologically evolve into a homogeneous population while still carrying very divergent matrilines through different colonizations and subsequent population introgression. For example, before data on their mtDNA were available, the genetically distinct ABC bears were not a recognized subspecies, while clearly morphologically distinct bears from Kamchatka and Kodiak were recognized as subspecies, but are just branches inside the 3a subclade. The Syrian bear comprises three major mtDNA clades (and seven subclades), and the Siberian brown bear seems to be a mixture of two. The recognition of local populations in the Apennines (U. a. marsicanus) and the Pyrenees (U. a. pyrenaicus) cannot be supported based on current genetic evidence (Loy et al. 2008; Colangelo et al. 2012). Paternal phylogeography is a new genetic approach that could be applied to this species, which is based on SNPs and short tandem repeats (STRs) located in the non-recombining fragment of the Y-chromosome (Bidon et al. 2014; Hirata et al. 2017). However, there are currently insufficient data for a comprehensive coverage of the brown bear’s widespread range to understand its geographical structure.

Population Genetic Assessment and Management of Bear Populations for Conservation Conservation genetics based on non-invasive genetic sampling has revolutionized how individuals, groups, and populations of bears are characterized and monitored (Waits et al. 1999; Schwartz et al. 2007). The majority of population genetic analyses in bears used microsatellites or STRs. However, markers based on genomic approaches, such as SNPs and genome sequencing of individuals, have been increasingly applied. Substantial improvements in methodology have led to the extensive use of non-invasive genetic sampling to

11

Systematics, Ecology, and Behavior

Figure 1.4 Bayesian phylogenetic tree of brown bears based on mitochondrial genomes (Anijalg et al. 2018). Numbers on nodes and after clades represent posterior mean estimates of the ages of different lineages, with numbers in parentheses representing 95% credibility intervals. Vertical gray lines represent long glacial periods. The recently described Himalayan brown bear clade is not shown (Lan et al. 2017), but is basal to clade 1, Western lineage.

identify and determine the numbers of bears in areas, countries, or even transborder regions (Bischof et al. 2016). Noninvasive samples are used for genetic assessments of populations, including estimation of genetic diversity and substructure, inbreeding, effective population size and thus population viability, and conservation status. Further, regular sampling allows development of databases containing many individual genotypes and also comprehensive ecological monitoring data of the same individuals over several years [e.g. Norway (see Figure 1.5) and the USA; see Chapter 13].

DNA-Based Monitoring of Bears Today, many techniques are used in DNA research and monitoring of bears. Different genetic markers and assessments may show different results owing to the resolution of genetic data. Wildlife researchers and managers must define a specific genetic sampling and monitoring scheme, and choose the genetic assessment method which fits their needs and delivers the

12

information needed to achieve management or research goals. Early studies of populations were based on high-quality DNA samples from blood or tissues collected from captured or dead bears. From such samples many genetic markers can be used to identify individuals, ascertain sex and assess genetic diversity, structure, kinship, and gene flow among individuals and populations, or to obtain long fragments of mtDNA to study phylogeny. When studying large, small or endangered bear populations, or when budgets are limited, non-invasive sampling (feces, hairs) provides opportunities to obtain genetic data without disturbing or even observing individuals and may be less expensive. The collected samples are usually of low-quality, somewhat degraded DNA, which constrains the number of markers that can be studied and the length of the studied DNA fragment, and therefore options for population genetic analyses (see Table 1.3). However, when samples are successfully genotyped, individual identification allows estimation of home ranges, population size through capture–mark–recapture

Systematics, Evolution, and Genetics of Bears

Figure 1.5 The Norwegian Large Predator Monitoring Program operated by Rovdata. Rovdata is responsible for the analyses and reports the results and monitoring data on brown bears in the country. Roughly 1000 non-invasive samples (mainly feces and hairs) are collected annually to assess the number of bears to guide the management by estimating the annual reproductive output (Bischof & Swenson 2012). Results of individual identifications, and other data, are accessible to the public at www.rovbase.no. From 2016 to 2019, for example, 2867 feces samples were collected and analyzed: successfully genotyped samples (n = 1770; black), negative samples (n = 1097; gray); figure by Kopatz et al. (in preparation, modified).

13

Systematics, Ecology, and Behavior Table 1.3 Applications for DNA-based studies of bears depending on the type of sample obtained. x, commonly observed in the literature; (x), observed in recent studies or potential application.

Species (mitochondrial DNA)

Individual (nuclear DNA)

Tissue samples

Fecal samples/hairs

Ancient DNA samples

eDNA samples/ (mixed genetic material)

Species identification

x

x

x

x

Lineage identification

x

x

x

Interspecific phylogeny

x

Genotyping

x

x

(x)

Sex identification

x

x

(x)

Genetic diversity

x

x

Abundance estimates

x

x

Structure/gene flow

x

(x)

Kinship analyses

x

(x)

methods, or population trends of identified groups or subpopulations based on genetic parameters (e.g. Kendall et al. 2009). Importantly for isolated and endangered species and populations, such genotyping data can also be used to identify MUs or ESUs. In scenarios where collection of tissues, feces or hairs is impossible, species or even individuals can be identified from environmental DNA (eDNA) samples, i.e. biological material collected from substrates such as soil, water, or snow that bears have encountered. Several recent studies applying eDNA technologies in various conditions and environments highlight the potential for this technology. For example, Wheat et al. (2016) collected saliva swabs from salmon carcasses and showed that this is a more cost-effective, rapid method to identify individual bears than the use of scats. Von Duyke et al. (2017) simulated fresh tracks by pressing the foot of a subsistence-harvested polar bear into snow. The genotype from the snow sample, based on next-generation sequencing (De Barba et al. 2017), matched the genotype obtained from this harvested bear. Further technical developments might allow researchers to obtain bear DNA from tracks in the field, from different substrates (such as soil around the den), or indirectly through invertebrate- or parasite-derived DNA (iDNA; Abrams et al. 2019). The possibility of identifying individuals with these new approaches could revolutionize the monitoring and management of bear populations.

Population Genetic and Genomic Assessments Genetic diversity provides a metric of genetic health, an evaluation of historical reductions in population size, and insights into the potential for long-term viability through adaptive capacity. Genetic diversity varies considerably across bear populations and is affected by the number and variability of loci along with sample

14

(x)

size. For example, in brown bears, heterozygosity ranges from 0.27 to 0.81 (Skrbinšek et al. 2012). As expected, low diversity occurs in small and isolated populations (e.g. brown bears in the Gobi Desert, Cantabrian Mountains, and Kodiak Island; Andean bears in Ecuador). Theoretical concepts and methodology to estimate effective population sizes in natural populations have been evaluated (Skrbinšek et al. 2012; Kamath et al. 2015; Kopatz et al. 2017) to determine whether they would be practical to assess and track the number of individuals effectively contributing to populations (Pierson et al. 2018). A study on a population of grizzly bears in Greater Yellowstone Ecosystem, USA demonstrated that numbers of bears and effective population size have increased, despite relatively low and stable genetic diversity (Kamath et al. 2015). While limited evidence for deleterious inbreeding effects exists in natural populations (but see Dunbar et al., 1996), decreased litter sizes and albinism have been demonstrated in captive brown bears (Laikre et al. 1996). Similarly, male Scandinavian brown bears with high interrelatedness had lower yearly reproductive success (Zedrosser et al. 2007), and, in the Apennine mountains, alleles predicted to be deleterious were fixed in brown bears, including several associated with health problems in humans (Benazzo et al. 2017). Low diversity has also been demonstrated in the captive-bred giant panda population (Shen et al. 2009) and genetic bottlenecks have been illustrated for American black bears (Murphy et al. 2017). However, genetic diversity can increase rapidly when populations and genes mix (e.g. Mikle et al. 2016). The importance of isolation has led many studies to assess population structure, often in combination with assignment tests and coalescent approaches, to evaluate connectivity between subpopulations. For example, Dutta et al. (2015) identified migrants between populations of Indian sloth bears.

Systematics, Evolution, and Genetics of Bears

Many studies use landscape–genetic methods to evaluate resistance of various landscape features, with the aim of managing landscapes to maintain or enhance connectivity between separated subpopulations. Draheim et al. (2018) found that American black bears showed a stronger signal of isolation by resistance over time as forest fragmentation increased. Parentage approaches have been used to evaluate movement (e.g. Graves et al. 2014), as well as to understand the impact of interactions of behavior and relatedness on fitness, through mechanisms such as multiple paternity, sexually selected infanticide, and site fidelity (e.g. Norman et al. 2017). Analyses of population structure are also used in forensics to identify the source of illegally killed individuals or bear body parts (Frosch et al. 2014). With the development of genomics, new approaches to diversity, inbreeding, and other questions are becoming more common, as well as the ability to evaluate the role of selection and adaptive capacity in population survival and evolution. Genetic information enables studies on relatedness and kinship. For example, Mikle et al. (2016) used a pedigree approach to document mechanisms behind changes in diversity, where initial low diversity in a population resulted from a few reproductively dominant individuals, but increases in migration between subpopulations over time led directly to increased diversity. Norman et al. (2017) developed a method to identify fine-scale population genetic structure, which is sensitive to spatial and temporal changes in the landscape.

Conclusions Despite the few extant ursid species, the evolution of bears has been relatively well studied. This is partly because of their large size, so they are common in the fossil record, but also because they compete for space and resources with people, and have been and continue to be targets for direct exploitation, including entertainment, sport hunting, and to provide fur and other tissues for human use. In recent years genetics have provided a

better understanding of relationships among bear species, which are complicated owing to ancient hybridizations, and with other carnivorans, although some uncertainties remain. Genetics have also improved our knowledge of variation within species, some of which, such as the brown bear, have very extensive geographical ranges, so that species’ histories and colonizations can be resconstructed in deeper time by comparison with the fossil record. Genetics are vital for the future successful conservation and management of bear populations, especially with the advent of new techniques in recent years. Whereas the population genetics of the European brown and grizzly bears, as well as the American black bear to some extent, are quite well studied in terms of population genetics, there are few genetic assessments for other bear species, e.g. sun bears, sloth bears, and Asian black bears. Increased human expansion and a changing climate will cause further fragmentation, stress populations, and endanger bear species across the world, which will challenge current conservation, management, and research. Therefore, the genetic assessment of bear populations and their status and connectivity will increase in importance. Attempts to rewild certain areas may include the release of bears, which should incorporate genetic monitoring programs. Bears will continue to play an important role in genetic research because they (especially brown bears) are ecologically among the most intensively studied species, which enables comprehensive studies on the interplay of wildlife ecology and genomics. Specifically, we expect future genetic studies to investigate further adaptation and natural selection of bears, inbreeding, and evaluate circumstances for which effective population size might be appropriate for tracking abundance of bears. As new tools that use genetic information for conservation questions emerge, and sampling, genotyping, and sequencing techniques improve and become cheaper, genetics will become even more essential for bear research, conservation, and management.

Abella, J., Alba, D. M., Robles, J. M., et al. (2012). Kretzoiarctos gen. nov., the oldest member of the giant panda clade. PLoS ONE 7: e48985.

Anijalg, P., Ho, S. Y. W., Davison, J., et al. (2018). Large-scale migrations of brown bears in Eurasia and to North America during the Late Pleistocene. Journal of Biogeography 45: 394–405. https:// doi.org/10.1111/jbi.13126

Abrams, J. F., Hoerig, L., Brozovic, R., et al. (2019). Shifting up a gear with iDNA: from mammal detection events to standardized surveys. Journal of Applied Ecology 56: 1637–1648. https://doi.org/10 .1111/1365-2664.13411

Ashrafzadeh, M. R., Kaboli, M. & Naghavi, M. R. (2016). Mitochondrial DNA analysis of Iranian brown bears (Ursus arctos) reveals new phylogeographic lineage. Mammalian Biology-Zeitschrift für Säugetierkunde 81(1): 1–9.

Ameghino, F. (1902). L’age des formations sédimentaires de Patagonie. Anales de la Sociedad Científica Argentina 54: 220–249.

Bapteste, E., van Iersel, L., Janke, A., et al. (2013). Networks: expanding evolutionary thinking. Trends in Genetics 29: 439–441.

Ameghino, F. (1903). L’age des formations sédimentaires de Patagonie. Buenos Aires: Imprimerie Coni Frères.

Barlow, A., Cahill, J. A., Hartmann, S., et al. (2018). Partial genomic survival of

References

cave bears in living brown bears. Nature Ecology & Evolution 2: 1563–1570. Barlow, A., Sheng, G. L., Lai, X. L., Hofreiter, M. & Paijmans, J. L. (2019). Once lost, twice found: combined analysis of ancient giant panda sequences characterises extinct clade. Journal of Biogeography 46: 251–253. Barnes, G. L. (1999). The rise of civilization in East Asia: The archaeology of China, Korea and Japan. London: Thames & Hudson. Barnes, I., Matheus, P., Shapiro, B., Jensen, D. & Cooper, A. (2002): Dynamics of Pleistocene population extinctions in Beringian brown bears. Science 295: 2267–2270.

15

Systematics, Ecology, and Behavior

Batsch, A. J. G. C. (1788). Versuch einer Anleitung, zur Kenntniß und Geschichte der Thiere und Mineralien, für akademische Vorlesungen entworfen, und mit den nöthigsten Abbildungen versehen. Jena: Akademische Buchhandlung. Benazzo, A., Trucchi, E., Cahill, J. A., et al. (2017). Survival and divergence in a small group: the extraordinary genomic history of the endangered Apennine brown bear stragglers. Proceedings of the National Academy of Sciences 114: E9589–E9597. Berta, A., Sumich, J. L. & Kovacs, K. M. (2015). Marine mammals: Evolutionary biology. London: Academic Press. Bidon T., Janke A., Fain S. R., et al. (2014). Brown and polar bear Y chromosomes reveal extensive male-biased gene flow within brother lineages. Molecular Biology and Evolution 31: 1353–1363. https://doi.org/10.1093/ molbev/msu109 Bischof, R. & Swenson, J. E. (2012). Linking noninvasive genetic sampling and traditional monitoring to aid management of a trans-border carnivore population. Ecological Applications 22(1): 361–373. Bischof, R., Brøseth, H. & Gimenez, O. (2016). Wildlife in a politically divided world: insularism inflates estimates of brown bear abundance. Conservation Letters 9: 122–130. de Bonis, L. (2013). Ursidae (Mammalia, Carnivora) from the Late Oligocene of the “Phosphorites du Quercy” (France) and a reappraisal of the genus Cephalogale Geoffroy, 1862. Geodiversitas 35: 787–814. de Bonis, L., Abella, J., Merceron, G. & Begun, D. R. (2017). A new late Miocene ailuropodine (giant panda) from Rudabánya (north-central Hungary). Geobios 50: 413–421. Bowdich, T. E. (1821). An analysis of the natural classifications of Mammalia, for the use of students and travellers. Paris: J. Smith. Bryant, H. N. (1996). Explicitness, stability, and universality in the phylogenetic definition and usage of taxon names: a case study of the phylogenetic taxonomy of the Carnivora (Mammalia). Systematic Biology 45(2): 174–189. Cahill, J. A., Green, R. E., Fulton, T. L., et al. (2013). Genomic evidence for island population conversion resolves conflicting theories of polar bear evolution. PLoS Genetics 9(3): e1003345.

16

Cahill, J. A., Stirling, I., Kistler, L., et al. (2015). Genomic evidence of geographically widespread effect of gene flow from polar bears into brown bears. Molecular Ecology 24: 1205–1217. Cahill, J. A., Heintzman, P. D., Harris, K., et al. (2018). Genomic evidence of widespread admixture from polar bears into brown bears during the Last Ice Age. Molecular Biology and Evolution 35: 1120–1129. Campagna L., Van Coeverden de Groot, P. J., Saunders, B. L., et al. (2013). Extensive sampling of polar bears (Ursus maritimus) in the Northwest Passage (Canadian Arctic Archipelago) reveals population differentiation across multiple spatial and temporal scales. Ecology and Evolution 3: 3152–3165. Chen, Y.-Y., Zhu, Y., Wan, Q.-H., et al. (2013). Patterns of adaptive and neutral diversity identify the Xiaoxiangling Mountains as a refuge for the giant panda. PLoS ONE 8(7): e70229. Chorn, J. & Hoffmann, R. S. (1978). Ailuropoda melanoleuca. Mammalian Species 110: 1–6. Çilingir, F. G., Akın Pekşen, Ç., Ambarlı, H., Beerli, P. & Bilgin, C. C. (2016). Exceptional maternal lineage diversity in brown bears (Ursus arctos) from Turkey. Zoological Journal of the Linnean Society 176: 463–477. Colangelo, P., Loy, A., Huber, D., et al. (2012). Cranial distinctiveness in the Apennine brown bear: genetic drift effect or ecophenotypic adaptation? Biological Journal of the Linnean Society 107: 15–26. Corbet, G. B. & Hill, J. E. (1992). The mammals of the Indomalayan region: A systematic review. London: Natural History Museum Publications and Oxford: Oxford University Press. Crompton, A. E., Obbard, M. E., Petersen, S. D. & Wilson, P. J. (2008). Population genetic structure in polar bears (Ursus maritimus) from Hudson Bay, Canada: implications of future climate change. Biological Conservation 141: 2528–2539.

coloriées, dessinées d’après des animaux vivans, pt. 3, vol. 5(5), 2 pp. Paris: A. Belin. Cuvier, G. (1823). Des ossemens d’ours. In Recherches sur les ossemens fossiles, où l’on rétablit les caractères de plusieurs animaux dont les révolutions du globe ont détruit les espèces, vol. 4(3), pp. 311–380. Paris: G. Dufour et E. d’Ocagne. David, A. (1869). Voyage en Chine, troisième partie. Nouvelles archives du Muséum d’histoire naturelle de Paris 5: 3–13. Davison, J., Ho, S. Y., Bray, S. C., et al. (2011). Late-Quaternary biogeographic scenarios for the brown bear (Ursus arctos), a wild mammal model species. Quaternary Science Reviews 30: 418–430. De Barba, M., Miquel, C., Lobréaux, S., et al. (2017). High-throughput microsatellite genotyping in ecology: Improved accuracy, efficiency, standardization and success with low-quantity and degraded DNA. Molecular Ecology Resources 17: 492–507. Doronina, L., Churakov, G., Shi, J., et al. (2015). Exploring massive incomplete lineage sorting in arctoids (Laurasiatheria, Carnivora). Molecular Biology and Evolution 32: 3194–3204. Draheim, H. M., Moore, J. A., Fortin, M. J. & Scribner, K. T. (2018). Beyond the snapshot: landscape genetic analysis of time series data reveal responses of American black bears to landscape change. Evolutionary Applications 11: 1219–1230. Dunbar, M. R., Cunningham, M. W., Wooding, J. B. & Roth, R. P. (1996). Cryptorchidism and delayed testicular descent in Florida black bears. Journal of Wildlife Diseases 32: 661–664. Durner, G. M., Laidre, K. L. & York, G. S. (Eds.) (2018). Polar bears: Proceedings of the 18th Working Meeting of the IUCN/ SSC Polar Bear Specialist Group, 7–11 June 2016, Anchorage, Alaska. Gland, Switzerland and Cambridge, UK: IUCN.

Cronin, M. A., Amstrup, S. C., Garner, G. W. & Vyse, E. R. (1991). Interspecific and intraspecific mitochondrial DNA variation in North American bears (Ursus). Canadian Journal of Zoology 69: 2985–2992.

Dutta, T., Sharma, S., Maldonado, J. E., Panwar, H. S. & Seidensticker, J. (2015). Genetic variation, structure, and gene flow in a sloth bear (Melursus ursinus) meta-population in the Satpura-Maikal landscape of Central India. PLoS ONE 10(5): e0123384.

Cuvier, F. (1825). Ours des Cordilères du Chili. In: Geoffroy-Saint-Hilaire, É. & Cuvier, F. (Eds.) Histoire naturelle des mammifères, avec des figures originales,

Edwards C. J., Suchard, M. A, Lemey, P., et al. (2011). Ancient hybridization and an Irish origin for the modern polar bear matriline. Current Biology 21(15): 1–8.

Systematics, Evolution, and Genetics of Bears

Eizirik, E., Murphy, W. J., Koepfli, K. P., et al. (2010). Pattern and timing of diversification of the mammalian order Carnivora inferred from multiple nuclear gene sequences. Molecular Phylogenetics and Evolution 56: 49–63.

Goswami, A. (2010). Introduction to Carnivora. In Goswami, A. & Friscia, A. (Eds.), Carnivoran evolution. New views on phylogeny, form and function, pp. 1–24. Cambridge: Cambridge University Press.

Ellegren, H., Primmer, C. R. & Sheldon, B. C. (1995). Microsatellite evolution: directionality or bias? Nature Genetics 11: 360–362.

Graves, T., Chandler, R. B., Royle, J. A., Beier, P. & Kendall, K. C. (2014). Estimating landscape resistance to dispersal. Landscape Ecology 29: 1201–1211.

Ellerman, J. R. & Morrison-Scott, T. C. S. (1951). Checklist of Palaearctic and Indian mammals 1758–1946. London: Trustees of the British Museum.

Grevé, C. (1894). Die geographische Verbreitung der jetzt lebenden Raubthiere. Kaiserlich LeopoldinischCarolinische Deutsche Akademie der Naturforscher. Nova Acta 63(1): 7–280.

Erdbrink, D. P. (1953): A review of fossil and recent bears of the Old World with remarks on their phylogeny based on their dentition, 2 vols. Deventer, the Netherlands: Jan de Lange. Ersmark, E., Baryshnikov, G., Higham, T., et al. (2019). Genetic turnovers and northern survival during the last glacial maximum in European brown bears. Ecology and Evolution 9: 5981–5905. https://doi.org/10.1002/ece3.5172 Flower, W. H. (1869). On the value of the characters of the base of the cranium in the classification of the order Carnivora, and on the systematic position of Bassaris and other disputed forms. Proceedings of the Zoological Society of London 37(1): 4–37. Frosch, C., Dutsov, A., Zlatanova, D., et al. (2014). Noninvasive genetic assessment of brown bear population structure in Bulgarian mountain regions. Mammalian Biology 79: 268–276. Galbreath, G. J., Hunt, M., Clements, T. & Waits L. P. (2008). An apparent hybrid wild bear from Cambodia. Ursus 19: 85–86. García-Rangel, S. (2012). Andean bear Tremarctos ornatus natural history and conservation. Mammal Review 42: 85–119. García-Vázquez, A., Pinto Llona, A. C. & Grandal-d’Anglade, A. (2017). Postglacial colonization of Western Europe brown bears from a cryptic Atlantic refugium out of the Iberian Peninsula. Historical Biology XX: 1–13. Gervais, P. (1855). Histoire naturelle des mammifères avec l’indication de leurs moeurs et de leurs rapports avec les arts, le commerce et l’agriculture. Paris: L. Curmer. Ginsburg, L. & Morales, J. (1998). Les Hemicyoninae (Ursidae, Carnivora, Mammalia) et les formes apparentées du Miocène inférieur et moyen d’Europe occidentale. Annales de Paléontologie 84: 71–123.

Hailer, F., Kutschera, V. E., Hallström, B. M., et al. (2012). Nuclear genomic sequences reveal that polar bears are an old and distinct bear lineage. Science 336(6079): 344–347. Hall, E. R. (1981): The mammals of North America, Vol. II (2nd edn.). New York, NY: John Wiley. Hassanin, A. (2015). The role of Pleistocene glaciations in shaping the evolution of polar and brown bears. Evidence from a critical review of mitochondrial and nuclear genome analyses. Comptes rendus biologies 338: 494–501. Hendey, Q. B. (1980). Agriotherium (Mammalia, Ursidae) from Langebaanweg, South Africa, and relationships of the genus. Annals of the South African Museum 81: 1–109. Hirata, D., Mano, T., Abramov, A. V., et al. (2017). Paternal phylogeographic structure of the brown bear (Ursus arctos) in northeastern Asia and the effect of male-mediated gene flow to insular populations. Zoological Letters 3 (1): 21. Hofreiter, M., Serre, D., Rohland, N., et al. (2004). Lack of phylogeography in European mammals before the last glaciation. Proceedings of the National Academy of Sciences 101: 12,963–12,968. Hu, Y., Qi, D., Wang, H. & Wei, F. (2010a). Genetic evidence of recent population contraction in the southernmost population of giant pandas. Genetica 138: 1297–1306.

Hunt Jr., R. M. (1977). Basicranial anatomy of Cynelos Jourdan (Mammalia: Carnivora), an Aquitanian amphicyonid from the Allier Basin, France. Journal of Paleontology 51(4): 826–843. Hunt Jr., R. M. (1998). Ursidae. In Janis, C. M., Scott, K. M. & Jacobs, L. L. (Eds.), Evolution of Tertiary mammals of North America, pp. 174–189. Cambridge: Cambridge University Press. Hwang, D.-S., Ki, J.-S., Jeong, D.-H., et al. (2008). A comprehensive analysis of three Asiatic black bear mitochondrial genomes (subspecies ussuricus, formosanus and mupinensis), with emphasis on the complete mtDNA sequence of Ursus thibetanus ussuricus (Ursidae). Mitochondrial DNA 19: 418–429. Kadariya, R., Shimozuru, M., Maldonado, J. E., Moustafa, M. A. M., Sashika, M. & Tsubota, T. (2018). High genetic diversity and distinct ancient lineage of Asiatic black bears revealed by non-invasive surveys in the Annapurna Conservation Area, Nepal. PLoS ONE, 13(12): e0207662. Kamath, P. L., Haroldson, M. A., Luikart, G., et al. (2015). Multiple estimates of effective population size for monitoring a long-lived vertebrate: an application to Yellowstone grizzly bears. Molecular Ecology 24: 5507–5521. Kendall, K. C., Stetz, J. B., Boulanger, J., et al. (2009). Demography and genetic structure of a recovering grizzly bear population. The Journal of Wildlife Management 73(1): 3–16. Kim, Y. K., Hong, Y. J., Min, M. S., et al. (2011). Genetic status of Asiatic black bear (Ursus thibetanus) reintroduced into South Korea based on mitochondrial DNA and microsatellite loci analysis. Journal of Heredity 102: 165–174. Kitchener, A. C., Breitenmoser-Würsten, Ch., Eizirik, E., et al. (2017). A revised taxonomy of the Felidae. The final report of the Cat Classification Task Force of the IUCN/SSC Cat Specialist Group. Cat News Special Issue 11: 1–80.

Hu, Y., Zhan, X., Qi, D. & Wei, F. (2010b). Spatial genetic structure and dispersal of giant pandas on a mountain-range scale. Conservation Genetics 11: 2145–2155.

Koepfli, K.-P., Dragoo, J. W. & Wang, X. (2017). The evolutionary history and molecular systematics of the Musteloidea. In: Macdonald, D. W., Newman, C. & Harrington, L. A. (Eds.), Biology and conservation of the musteloids, pp. 92–128. Oxford: Oxford University Press.

Hunt Jr., R. M. (1974). The auditory bulla in Carnivora: an anatomical basis for reappraisal of carnivore evolution. Journal of Morphology 143(1): 21–75.

Kopatz, A., Eiken, H. G., Schregel, J., et al. (2017). Genetic substructure and admixture as important factors in linkage disequilibrium-based estimation of

17

Systematics, Ecology, and Behavior

effective number of breeders in recovering wildlife populations. Ecology and Evolution 7: 10,721–10,732. Koretsky, I. A., Barnes, L. G. & Rahmat, S. J. (2016). Re-evaluation of morphological characters questions current views of pinniped origins. Vestnik Zoologii 50: 327–354. Krause, J., Unger, T., Nocon, A., et al. (2008). Mitochondrial genomes reveal an explosive radiation of extinct and extant bears near the Miocene–Pliocene boundary. BMC Evolutionary Biology 8(1): 220. Kretzoi, M. (1943). Kochitis centennii n. g. n. sp. az egeresi felső oligocénből [in Hungarian]/Kochitis centenii n.g. n.sp., ein altertümlicher Creodonte aus dem Oberoligozän Siebenbürgens [in German]. Földtani Közlöny 73: 10–17 [Hungarian] / 190–195 [German]. Kumar, V., Lammers, F., Bidon, T., et al. (2017). The evolutionary history of bears is characterized by gene flow across species. Scientific Reports 7: 46487. Kurtén, B. & Anderson, E. (1980). Pleistocene mammals of North America. New York, NY: Columbia University Press. Kutschera, V. E., Bidon, T., Hailer, F., et al. (2014). Bears in a forest of gene trees: phylogenetic inference is complicated by incomplete lineage sorting and gene flow. Molecular Biology and Evolution 31: 2004–2017. Kutschera, V. E., Frosch, C., Janke, A., et al. (2016). High genetic variability of vagrant polar bears illustrates importance of population connectivity in fragmented sea ice habitats. Animal Conservation 19: 337–349.

18

adaptation in polar bears. Cell 157: 785–794. Loy, A., Genov, P., Galfo, M., Jacobone, M. G. & Vigna Taglianti, A. (2008). Cranial morphometrics of the Apennine brown bear (Ursus arctos marsicanus) and preliminary notes on the relationships with other southern European populations. Italian Journal of Zoology 75: 67–75. Luan, P. T., Ryder, O. A., Davis, H., Zhang, Y. P. & Yu, L. (2013). Incorporating indels as phylogenetic characters: impact for interfamilial relationships within Arctoidea (Mammalia: Carnivora). Molecular Phylogenetics and Evolution 66: 748–756.

Mitchell, K. J., Bray, S. C., Bover, P., et al. (2016). Ancient mitochondrial DNA reveals convergent evolution of giant short-faced bears (Tremarctinae) in North and South America. Biology Letters 12(4): 20160062. Murphy, S. M., Augustine, B. C., Ulrey, W. A., et al. (2017). Consequences of severe habitat fragmentation on density, genetics, and spatial capture-recapture analysis of a small bear population. PLoS ONE 12(7): e0181849. Murtskhvaladze M., Gavashelishvili A. & Tarkhnishvili D. (2010). Geographic and genetic boundaries of brown bear (Ursus arctos) population in the Caucasus. Molecular Ecology 19: 1829–1841.

Malenfant, R. M., Davis, C. S., Cullingham, C. I. & Coltman, D. W. (2016). Circumpolar genetic structure and recent gene flow of polar bears: a reanalysis. PLoS ONE 11(3): e0148967.

Norman, A. J., Stronen, A. V., Fuglstad, G. A., et al. (2017). Landscape relatedness: detecting contemporary fine-scale spatial structure in wild populations. Landscape Ecology 32: 181–194.

Margari, V., Skinner, L. C., Tzedakis, P. C., et al. (2010). The nature of millennialscale climate variability during the past two glacial periods. Nature Geoscience 3: 127–131.

Nyakatura, K. & Bininda-Emonds, O. R. (2012). Updating the evolutionary history of Carnivora (Mammalia): a new specieslevel supertree complete with divergence time estimates. BMC Biology 10(1): 12.

Meijaard, E. (2004): Craniometric differences among Malayan sun bears (Ursus malayanus); evolutionary and taxonomic implications. The Raffles Bulletin of Zoology 52: 665–672.

Ohnishi, N., Uno, R., Ishibashi, Y., Tamate, H. B. & Oi, T. (2009). The influence of climatic oscillations during the Quaternary Era on the genetic structure of Asian black bears in Japan. Heredity 102: 579.

Merriam J. C. & Stock C. (1925). Relationships and structure of the shortfaced bear, Arctotherium, from the Pleistocene of California. Contributions to Palaeontology, Carnegie Institution of Washington 347: 1–35.

Onuma, M., Suzuki, M. & Ohtaishi, N. (2006). Possible conservation units of the sun bear (Helarctos malayanus) in Sarawak based on variation of mtDNA control region. Japanese Journal of Veterinary Research 54: 135–139.

Laikre, L., Andrén, R., Larsson, H. O. & Ryman, N. (1996). Inbreeding depression in brown bear Ursus arctos. Biological Conservation 76: 69–72.

Meyer, F. A. (1793). Systematischsummarische Uebersicht der neuesten zoologischen Entdeckungen in Neuholland und Afrika. Leipzig: Dykischen Buchhandlung.

Lan, T., Gill, S., Bellemain, E., et al. (2017). Evolutionary history of enigmatic bears in the Tibetan Plateau–Himalaya region and the identity of the yeti. Proceedings of the Royal Society B: Biological Sciences 284 (1868): 20171804.

Mikle, N., Graves, T. A., Kovach, R., Kendall, K. C. & Macleod, A. C. (2016). Demographic mechanisms underpinning genetic assimilation of remnant groups of a large carnivore. Proceedings of the Royal Society of London B 283(1839): 20161467.

Linnaeus, C. V. (1758). Systema Naturae per regna tria naturae. Secundum classes, ordines, genera, species, cum characteribus, differentiis, synonymis, locis, 10th edition, Vol. 1. Stockholm: Lars Salvius.

Miller, W., Schuster, S. C., Welch, A. J., et al. (2012). Polar and brown bear genomes reveal ancient admixture and demographic footprints of past plimate change. Proceedings of the National Academy of Sciences 109: E2382–E2390.

Pallas, P. S. (1780). Spicilegia zoologica: quibus novae imprimis et obscurae animalium species iconibus, descriptionibus atque commentariis illustrantur, fascicle 14, pp. 3–24. Berlin: Joachim Paul.

Liu, S., Lorenzen, E. D., Fumagalli, M., et al. (2014). Population genomics reveal recent speciation and rapid evolutionary

Milne-Edwards, A. (1870). Note sur quelques mammifères du Thibet oriental. Annales des Sciences Naturelles, Zoologie et Paléontologie, 5th series, 13, article 10.

Pamilo, P. & Nei, M. (1988). Relationships between gene trees and species trees. Molecular Biology and Evolution 5: 568–583.

Paetkau, D., Shields, G. F. & Strobeck, C. (1998). Gene flow between insular, coastal and interior populations of brown bears in Alaska. Molecular Ecology 7: 1283–1292. Pages, M., Calvignac, S., Klein, C., et al. (2008). Combined analysis of fourteen nuclear genes refines the Ursidae phylogeny. Molecular Phylogenetics and Evolution 47: 73–83.

Systematics, Evolution, and Genetics of Bears

Phipps, C. J. (1774). A voyage towards the North Pole undertaken by His Majesty’s command. London: J. Nourse.

Ruiz-García, M., Orozco-terWengel, P., Payán, E. & Castellanos, A. (2003). Genética de Poblaciones molecular aplicada al estudio de dos grandes carnívoros (Tremarctos ornatus – Oso andino, Panthera onca – jaguar): lecciones de conservación. Boletín de la Real Sociedad Española de Historia Natural 98: 135–158.

Pierson, J. C., Graves, T. A., Banks, S. C., Kendall, K. C. & Lindenmayer, D. B. (2018). Relationship between effective and demographic population size in continuously distributed populations. Evolutionary Applications 11: 1162–1175.

Ruiz-García, M., Orozco-terWengel, P., Castellanos, A. & Arias, L. (2005). Microsatellite analysis of the spectacled bear (Tremarctos ornatus) across its range distribution. Genes and Genetics Systems 80: 57–69.

Puckett, E. E., Etter, P. D., Johnson, E. A. & Eggert, L. S. (2015). Phylogeographic analyses of American black bears (Ursus americanus) suggest four glacial refugia and complex patterns of postglacial admixture. Molecular Biology and Evolution 32: 2338–2350.

Ruiz-García, M., Arias, J. Y., Castellanos, A., Kolter, L. & Shostell, J. M. (2020a). Molecular evolution (mitochondrial and nuclear microsatellites markers) in the Andean bear (Tremarctos ornatus; Ursidae, Carnivora): How many ESUs are there? In: Ortega, J. & Maldonado J. E. (Eds.), Mammalian Conservation Genetics (pp. 165–194). Berlin: Springer-Verlag.

Pelletier, A., Obbard, M. E., White, B. N., Doyle, C. & Kyle, C. J. (2011). Small-scale genetic structure of American black bears illustrates potential postglacial recolonization routes. Journal of Mammalogy 92: 629–644.

Raffles, T. S. (1821). Descriptive catalogue of a zoological collection, made on account of the Honourable East India Company, in the island of Sumatra and its vicinity, under the direction of Sir Thomas Stamford Raffles, Lieutenant-Governor of Fort Marlborough; with additional notices illustrative of the natural history of those countries. Transactions of the Linnean Society of London 13(1): 239–274. Ren, G. (2000). Decline of the mid- to late Holocene forests in China: climatic change or human impact? Journal of Quaternary Science 15: 273–281. Rose, K. D. (2006). The beginning of the age of mammals. Baltimore, MD: Johns Hopkins University Press.

Ruiz-García, M., Arias, J. Y., Restrepo, H., Cáceres-Martínez, C. & Shostell, J. M. (2020b). The genetic structure of the Andean bear (Tremarctos ornatus; Ursidae, Carnivora) in Colombia by means of mitochondrial and microsatellite markers. Journal of Mammalogy (in press). Ruiz-García, M., Castellanos, A., Arias, J.Y. & Shostell, J.M. (2020c). Genetics of the Andean bear (Tremarctos ornatus; Ursidae, Carnivora) in Ecuador: When the Andean Cordilleras are not an obstacle. Mitochondrial DNA Part A 31: 194–212.

management. Trends in Ecology & Evolution 22: 25–33. Shaw, G. (1790). The naturalist’s miscellany, or coloured figures of natural objects, vol. 2. London: Nodder & Co. Shen, F., Zhang, Z., He, W. E. I., et al. (2009). Microsatellite variability reveals the necessity for genetic input from wild giant pandas (Ailuropoda melanoleuca) into the captive population. Molecular Ecology 18: 1061–1070. Skrbinšek, T., Jelenčič, M., Waits, L., et al. (2012). Monitoring the effective population size of a brown bear (Ursus arctos) population using new singlesample approaches. Molecular Ecology 21: 862–875. Soibelzon, L. H. & Schubert, B. W. (2011). The largest known bear, Arctotherium angustidens, from the Early Pleistocene Pampean Region of Argentina: with a discussion of size and diet trends in bears. Journal of Paleontology 85: 69–75. Soibelzon, L. H., Tonni, E. P. & Bond, M. (2005). The fossil record of South American short-faced bears (Ursidae, Tremarctinae). Journal of South American Earth Sciences 20: 105–113. Talbot, S. L. & Shields, G. F. (1996). A phylogeny of the bears (Ursidae) inferred from complete sequences of three mitochondrial genes. Molecular Phylogenetics and Evolution 5: 567–575. Tedford, R. H. (1976). Relationships of pinnipeds to other carnivores (Mammalia). Systematic Zoology 25: 363–374. Thenius, E. (1989). Molekulare und adaptive Evolution, Kladistik und Stammesgeschichte1: Ergänzungen zu einer Arbeitshypothese. Journal of Zoological Systematics and Evolutionary Research 27: 94–105.

Rounds, R. C. (1987). Distribution and analysis of colourmorphs of the black bear (Ursus americanus). Journal of Biogeography 14: 521–538.

Salesa, M. J., Siliceo, G., Antón, M., et al. (2006). Anatomy of the “false thumb” of Tremarctos ornatus (Carnivora, Ursidae, Tremarctinae): phylogenetic and functional implications. Estudios Geológicos 62: 389–394.

Ruiz-García, M. (2007). Genética de Poblaciones: Teoría y aplicación a la conservación de mamíferos neotropicales (Oso andino y delfín rosado). Boletín de la Real Sociedad Española de Historia Natural 102: 99–126.

Salomashkina, V. V., Kholodova, M. V., Semenov, U. A., Muradov, A. S. & Malkhasyan, A. (2017). Genetic variability of brown bear (Ursus arctos L., 1758). Russian Journal of Genetics 53: 108–117.

Ruiz-García, M. (2013). The genetic demography history and phylogeography of the Andean bear (Tremarctos ornatus) by means of microsatellites and mtDNA markers. In: Ruiz-García, M. & Shostell, J.M. (Eds.), Molecular population genetics, evolutionary biology and conservation of Neotropical carnivores (pp. 129–158). New York, NY: Nova Science Publishers.

Schmidt-Kittler, N. (1987). The Carnivora (Fissipedia) from the lower Miocene of East Africa. Palaeontographica. Abteilung A, Paläozoologie, Stratigraphie 197: 85–126.

Valdiosera, C. E., García, N., Anderung, C., et al. (2007). Staying out in the cold: glacial refugia and mitochondrial DNA phylogeography in ancient European brown bears. Molecular Ecology 16: 5140–5148.

Schwartz, M. K., Luikart, G. & Waples, R. S. (2007). Genetic monitoring as a promising tool for conservation and

Viengkone, M., Derocher, A. E., Richardson, E. S., et al. (2016). Assessing polar bear (Ursus maritimus) population structure

Turner, H. N. (1848). Observations relating to some of the foramina at the base of the skull in Mammalia, and on the classification of the order Carnivora. Proceedings of the Zoological Society of London 16: 63–88.

19

Systematics, Ecology, and Behavior

in the Hudson Bay region using SNPs. Ecology and Evolution 6: 8474–8484. Von Duyke, A., Bellemain, E., Dejean, T., et al. (2017). The future is now; science in a spoonful of snow. Using eDNA to monitor a polar bear (Ursus maritimus) population in the Chukchi Sea. 22nd Biennial Conference on the Biology of Marine Mammals. Wagner, J. (2010). Pliocene to early Middle Pleistocene ursine bears in Europe: a taxonomic overview. Journal of the National Museum (Prague), Natural History Series 179(20): 197–215. Waits, L., Paetkau, D. & Strobeck, C. (1999). Genetics of the bears of the world. In: Servheen, C., Herrero, S. & Peyton, B. (Eds.) Bears. Status survey and conservation action plan, pp. 25–32. Bern and Cambridge: IUCN. Wan, Q. H., Fang, S. G., Wu, H. & Fujihara, T. (2003). Genetic differentiation and subspecies development of the giant panda as revealed by DNA fingerprinting. Electrophoresis 24: 1353–1359. Wan, Q.-H., Wu, H. & Fang, S.-G. (2005). A new subspecies of giant panda (Ailuropoda melanoleuca) from Shaanxi, China. Journal of Mammalogy 86: 397–402. Wheat, R. E., Allen, J. M., Miller, S. D., Wilmers, C. C. & Levi, T. (2016).

20

Environmental DNA from residual saliva for efficient noninvasive genetic monitoring of brown bears (Ursus arctos). PLoS ONE 11(11): e0165259. Wilson, D. E. (1976). Cranial variation in polar bears. In Pelton, M. R., Lentfer, J. W. & Folk, G. E. (Eds.) Bears: Their biology and management. A selection of papers from the third international conference on bears, pp. 447–453. Morges, Switzerland: IUCN. Wooding, S. & Ward, R. (1997). Phylogeography and Pleistocene evolution in the North American black bear. Molecular Biology and Evolution 14: 1096–1105. Wozencraft, W. C. (1989). The phylogeny of the recent Carnivora. In: Gittleman, J. L. (Ed.) Carnivore behaviour, ecology and evolution (pp. 495–535). London: Chapman and Hall. Wu, J., Kohno, N., Mano, S., et al. (2015). Phylogeographic and demographic analysis of the Asian black bear (Ursus thibetanus) based on mitochondrial DNA. PLoS ONE 10(9): e0136398. Yu, L., Li, Q.-W., Ryder, O. A. & Zhang, Y.-P. (2004). Phylogeny of the bears (Ursidae) based on nuclear and mitochondrial genes. Molecular Phylogenetics and Evolution 32: 480–494.

Yu, L., Li, Y. W., Ryder, O. A. & Zhang, Y. P. (2007). Analysis of complete mitochondrial genome sequences increases phylogenetic resolution of bears (Ursidae), a mammalian family that experienced rapid speciation. BMC Evolutionary Biology 7(1): 198. Zedrosser, A., Støen, O. G., Sæbø, S. & Swenson, J. E. (2007). Should I stay or should I go? Natal dispersal in the brown bear. Animal Behaviour 74: 369–376. Zhang, B., Li, M., Zhang, Z., et al. (2007). Genetic viability and population history of the giant panda, putting an end to the “evolutionary dead end”? Molecular Biology and Evolution 24: 1801–1810. Zhang, Y. P. & Ryder, O. A. (1993). Mitochondrial DNA sequence evolution in the Arctoidea. Proceedings of the National Academy of Sciences of the United States of America 90: 9557–9561. Zhao, S., Zheng, P., Dong, S., et al. (2013). Whole-genome sequencing of giant pandas provides insights into demographic history and local adaptation. Nature Genetics 45(1): 67. Zuckerkandl, E. & Pauling L. (1962). Molecular disease, evolution and genetic heterogeneity. In: Horizons in biochemistry (pp. 189–225). New York, NY: Academic Press.

Chapter

2

Mating Strategies Sam M.J.G. Steyaert, Andreas Zedrosser, Ronald R. Swaisgood, Eva Filipczykova, Brian Crudge, Trishna Dutta, Sandeep Sharma, Shyamala Ratnayeke, Shinsuke Koike, Martin Leclerc, Andrew E. Derocher, Melanie Clapham, Thomas Spady, Bruce McLellan, Andre´s Ordiz, Alberto Ferna´ndez-Gil, Miguel Delibes, and Jon E. Swenson

Introduction The mating system and mating strategies of a species refer to the behavioral strategies used to obtain reproductive partners and ensure reproductive success (Emlen & Oring 1977). Common determining factors of mating systems and strategies are the manner of mate acquisition, the number of mates obtained by an individual, as well as the absence or presence and duration of parental care (Emlen & Oring 1977). In mammals, the energetic investments in gametes and rearing offspring are typically larger for females than for males (Trivers 1972; Andersson 1994). Mate selection is thus a much more important decision for females than for the rather indiscriminate males (Trivers 1972). This dichotomy results in sexual selection, which in turn is determined by male–male competition for access to females, as well as female mate choice (Andersson 1994). Because receptive females generally are considered as the limiting resource in reproduction, males face intrasexual competition for mates (Trivers 1972). In a multitude of mammalian species, including bears, this has resulted in pronounced sexual size dimorphism and polygamous mating systems (Andersson 1994). Despite common characteristics (e.g. sexual size dimorphism, polygamy), variation in mating systems and strategies occurs among bear populations and species.

Mating Systems of Bears of the World Giant Panda The giant panda’s mating system is best classified as scramble competition polygyny. Males occupy home ranges that are not mutually exclusive, overlapping with other male and female ranges. Average home range size is ~7 km2, and male ranges are about 1.5 times larger than female ranges (Connor et al. 2016). Females enter estrus asynchronously during the welldefined mating season from March to May (Schaller et al. 1985; Nie et al. 2012b; Pan 2014). When a female enters estrus, she displays increased movement and scent marking with urine and anogenital gland secretions, which serve to attract multiple males (Schaller et al. 1985). Up to five or six males congregate around the estrous female and compete vigorously for access to the female, and injuries are common. Males appear to gain access to females through a combination of their ability to locate

females (scramble competition) and direct physical competition with other males, with the largest, strongest males securing mating access (contest competition). Once relative competitive ability has been established among males through indirect assessment and escalated fighting where necessary, aggression among males subsides, but all males appear to remain with the aggregation until after the female has mated (Nie et al. 2012b). Females typically mate within their home range, although they occasionally move long distances outside of their home range during the mating season (Zhang et al. 2014). When a female enters the fertile period and is accompanied by several males, she typically climbs a tree and observes the competing males on the ground beneath her. Although difficult to determine, these observations may be used in mate selection. Consistent with this mating strategy that emphasizes male competitive ability, pandas show body size sexual dimorphism, with males weighing ~15% more than females. Given the importance of male competitive ability in determining reproduction outcomes, it is somewhat surprising that size dimorphism is not greater. Larger male body size may be constrained by the energetic requirements of participating in scramble and direct competition in a species that is highly energetically challenged (Nie et al. 2015). Pandas are obligate bamboo specialists, and bamboo comprises >98% of their diet (see also Chapter 6). Although the panda gut contains microbial symbionts that help break down cellulose and extract more nutritional value from bamboo, bamboo remains a very lowenergy food source (Zhu et al. 2011). Thus, it is plausible that the panda’s mating strategy is constrained energetically. Indeed, wild male pandas do not appear able to maintain energetically expensive elevated androgens during the mating season, as observed in most seasonally breeding mammals, and instead show elevated androgen only when challenged by the presence of an estrous female (Nie et al. 2012a). Females typically mate with only one male during a given mating season (Schaller et al. 1985; Nie et al. 2012b; Pan 2014), male pandas participate in several mating aggregations each season, and the most competitive males may secure copulations with multiple females. The interbirth interval for females is 1–3 years and the female fertile period lasts 2–3 days, producing a highly skewed operational sex ratio, the driving force behind male–male

21

Systematics, Ecology, and Behavior

competition. Although signs of intrasexual selection are easily observed, there is anecdotal evidence of female choice in the wild (Pan 2014). In captivity, studies have shown that both male and female mate choice occurs, and the opportunity to choose one’s own mate has substantial consequences for reproductive success in the form of copulation and birth rates (Martin-Wintle et al. 2015; see also Chapter 23). Personality – repeatable individual behavioral differences across time and situations – is one of the factors used in mate choice in pandas (Martin-Wintle et al. 2017). Reproductive maturity is reached in males and females at approximately 4.5 years of age. First reproduction for females typically occurs at age 6, and females can produce 7–8 cubs during their lifetime. The age of first reproduction in males is not known, but is presumed to be 6–7 years, when they gain sufficient size and competitive ability. The physiological parameters of the female reproductive cycle have been well documented, including temporal patterns of estrogens, progestagens, corticoids, and other hormones that characterize the estrous cycle, ovulation, pregnancy, and birth (Chapter 6). Pandas are spontaneous ovulators and, like other ursids, experience a variable period of embryonic diapause lasting from 50 to 108 days (Spady et al. 2007). Female litter size is 1–2 cubs (rarely 3), with twins occurring about half of the time (Zhu et al. 2001). Pandas also appear to engage in obligate litter reduction, abandoning one cub when there are twins, although there are rare reported cases of females rearing twins in the wild and captivity (Schaller et al. 1985). Pandas give birth to highly vulnerable offspring weighing about 1/900 the weight of the mother, making them the most altricial of all eutherian mammals (Gittleman 1994). This relatively small litter size and birth weight compared to other ursids is believed to be the consequence of energy limitation stemming from the panda’s low-energy bamboo diet. Giant panda females gained an unfavorable reputation as poor mothers from early challenges in captive breeding programs, but it now seems clear that panda maternal care is highly specialized and competent (Zhu et al. 2001; Pan 2014). Following birth, the female remains with the cub for two or more weeks before leaving briefly to drink, urinate, and defecate. Unlike many bear species, panda mothers actively hold their single cub on the ventrum, and remain attentive to infant vocalizations and postural adjustments. As the cub ages and gains independence, the mother begins to leave the cub behind for several hours while she forages. Older cubs typically spend this time safely sequestered in a tree. Cubs remain dependent on the mother for 1.5 years, occasionally 2.5 years. Unlike most other mammals, pandas exhibit female-biased dispersal, with subadult females dispersing out of their natal area (Zhang et al. 2014).

Andean Bear Andean bears are sexually size dimorphic (Garcia-Rangel 2012). Adult males measure 1.5–2 m and weigh 100–175 kg (Peyton 1980; Garshelis 2009; see Chapter 7), and females

22

reach about 2/3 of the adult male size (Garshelis 2009; Hunter & Barrett 2019). Females typically reach sexual maturity between 4 and 7 years old, and males usually mature when they are about 5 years old (Rosenthal 1999; Garshelis 2009; Garcia-Rangel 2012). Reproductive seasonality in Andean bears has mainly been studied in captivity (Spady et al. 2007). Females are polyestrous and reproductive cycles take at least 9 months, albeit three cycles sometimes follow each other within 24 months (Rosenthal 1999; Enciso & Guimarães 2013). The further away from the Equator, the stronger Andean bears exhibit seasonality in the timing of the breeding season (Appleton et al. 2018). Generally, mating in wild Andean bears occurs throughout the year, but peaks during the fruiting season (March–October; Garcia-Rangel 2012). However, Appleton et al. (2018) showed that Andean bears in the dry forests of Peru regularly mate in December and January. In zoos, mating occurs from February to September, depending on the latitude and photoperiod (Spady et al. 2007). Appleton et al. (2018) suggested that Andean bears display phenotypic plasticity in the timing of reproduction, possibly related to climate and food availability. The breeding season of Andean bears is relatively short. Appleton et al. (2018) reported that the breeding season in the wild generally lasts for about 6 weeks. Bears in captivity typically mate during a period of one to two weeks (Garcia-Rangel 2012). Captive females are in estrus for 3–14 days, which can be recognized by a swollen vulva. On average 2–8 copulations occur per day, each taking about 10–45 minutes. During courtship, couples play nonaggressive games accompanied by various rumbles. Andean bears communicate their reproductive and dominance status in the wild through scent marking (Filipczyková et al. 2017; Kleiner et al. 2018), although more evidence is needed about the functional significance of their marking behavior. Female Andean bears have delayed implantation. Gestation time is short and difficult to estimate, due to embryonic diapause (Garcia-Rangel 2012). In captivity, gestation takes from 160 to 210 days, but sometimes as long as 250 days (Rosenthal & Haggerty 1989). It appears that parturition, as with mating, occurs seasonally (Garcia-Rangel 2012; Appleton et al. 2018), with the majority of captive births occurring during autumn or winter. In the tropics, most births occur during the autumn, whereas most births outside the tropics occur during winter. Seasonality in births of wild Andean bears has been recorded in Peru (August–October) and in Ecuador (April), and might be related to the climate and plant phenology (Appleton et al. 2018; Castellanos et al. 2018). Females usually give birth for the first time when they are about 5 years old. As for other ursids and mammals in general, litter size depends on the body condition of the female and on food availability (Peyton 1980). The litter size ranges from one to four, but two cubs is most common (Rosenthal & Haggerty 1989). Birth occurs in a den and newborn cubs weigh approximately 275–380 g (Garcia-Rangel 2012). The mother and her dependent cubs stay in or in the immediate vicinity of the maternal den for

Mating Strategies

Figure 2.1 A female Andean bear with her cub, high up in a tree in Cali Zoo, Colombia. The visible nipples of the mother suggest that she is still nursing her cub, or had recently stopped nursing at that time of photographing. (Photo by E. Filipczykova.)

3–4 months (Castellanos et al. 2018). It has been reported that after this time, wild females with cubs change dens several times, possibly because of food availability and safety reasons. Mothers lactate for up to approximately 12 months, possibly even up to 14 months (Garcia-Rangel 2012; Castellanos et al. 2018; Figure 2.1; see Chapter 7). Cubs stay with their mothers on average for about two years, and females communicate with their offspring through a series of sounds. After weaning, captive females reportedly enter estrus after about 4–6 weeks. Annual survival is low (59–64%) for newborn captive Andean bear cubs (Garcia-Rangel 2012). Andean bears can coexist in multigenerational groups. In Northern Ecuador, for example, Castellanos and Vasquez (2017) reported that a single subadult female from a previous litter joined her mother, which had cubs-of-the-year. The subadult female interacted with both her mother and the cubs, and accompanied the cubs when the mother left to search for food. The subadult female stayed with the group until the mother separated from them. Later, the subadult female stayed within her mother’s home range and produced offspring of her own. Such observations suggest that Andean bear populations may have complex and kin-related social–spatial structures.

Sun Bear Sun bears are unique among ursids in several aspects of reproduction. Little is known about the sun bear’s reproductive behavior and social organization in the wild, but some insight on reproductive biology has been gained, predominantly from studies involving captive bears. There is limited sexual dimorphism in the body size of sun bears. In Borneo, male and female body mass averages about 40 and 25 kg, respectively. Healthy adults in captivity at a rescue facility in Cambodia (Free the Bears, https:// freethebears.org/) typically weigh about 70 kg for males and 60 kg for females. The weights of these bears revealed that adult females varied little seasonally, whereas males exhibited a

pronounced fluctuation, with a decline (ca. 12%) during the hot dry season (March–May; Crudge et al. 2019). Sun bears do not have a distinct breeding season (Schwarzenberger et al. 2004; Spady et al. 2007; Frederick et al. 2012; see also Chapter 8). Although they are capable of breeding throughout the year, more cubs are born during the wet season compared to the dry season in some parts of their range (Schwarzenberger et al. 2004). In a captive population, births occur during all months of the year, but least frequently during spring and most frequently during autumn and winter (Frederick et al. 2012). The absence of clear reproductive seasonality may be an adaptation for living in habitats with unpredictable food resources, like the populations that inhabit aseasonal evergreen rainforests in Peninsular Malaysia, Sumatra, and Borneo. These populations experience synchronized fruit masting events, followed by intermast periods with little fruiting. These large-scale fluctuations occur at intervals of 2–10 years. During extended periods of low fruit availability, loss of body mass and starvation occurs (Wong et al. 2005; Fredriksson & Wich 2006; Fredriksson et al. 2007). The unpredictability of fruit availability in such habitats may favor opportunistic, rather than seasonal, reproductive strategies (Frederick et al. 2012). With the exception of family groups, sun bears are solitary and associate only for mating (Ngoprasert et al. 2012). In captivity, sun bears have been shown to be spontaneous ovulators, but the presence of males had an influence on hormone metabolite concentrations and cytological profiles in females. The presence of males also coincided with an increased likelihood that females would cycle (Frederick et al. 2013). In a captive population, the age range for females giving birth was 4–27.8 years (mean = 10.5) and the reproductive age range for males was 4.2–26.6 years (mean = 11.0) (Frederick et al. 2013). Nonpregnant sun bears are polyestrous throughout the year, with up to three estrous cycles per year (Frederick et al. 2010). Like other bear species, sun bears routinely exhibit pseudopregnancy (Schwarzenberger et al. 2004; Frederick et al. 2010). Sun bears are capable of initiating estrus following neonatal mortality (Frederick et al. 2012). Gestation in sun bears is approximately 95–107 days, which is less variable than the other ursids (Schwarzenberger et al. 2004). Studies suggest that sun bears do not exhibit delayed implantation (see “Reproductive physiology”; Schwarzenberger et al. 2004; Frederick et al. 2012). Litters typically consist of one cub, or occasionally twins. In the wild, cubs will likely remain with their mother until fully grown at about 2 years of age (Schwarzenberger et al. 2004).

Sloth Bear Sloth bears exhibit size dimorphism with adult males weighing 80–150 kg and females weighing 60–100 kg (Prater 1965; Garshelis et al. 1999; Yoganand 2005). Home ranges of adult and subadult sloth bears overlap within and between sexes (Joshi et al. 1999; Yoganand 2005; Ratnayeke et al. 2007), in accordance with their polygamous mating system (see also Chapter 9).

23

Systematics, Ecology, and Behavior

Some of the best evidence of mating behavior in the wild comes from Chitwan National Park, Nepal, where behavioral observations of sloth bears were made from the backs of elephants (Laurie & Seidensticker 1977; Joshi et al. 1999). Males may make long excursions out of home range boundaries (>8 km) to mate. Females in estrus are accompanied by several males and may mate multiple times with certain males. Although dominance relationships among males may determine which male has priority of access to a receptive female, fights and loud vocalizations are a common feature of mating behavior in sloth bears (Laurie & Seidensticker 1977; Phillips 1984; Joshi et al. 1999). Joshi et al. (1999) report some evidence of female mate choice in such mating aggregations, including that of a female moving 2 km out of her home range to mate with a large male. Females may breed first when 4 years old, but do not necessarily produce cubs following their first breeding season. Pregnant females use dry caves among boulders and rock outcrops (Phillips 1984), or create dens by excavating holes in the banks of dry streambeds (Joshi et al. 1999). Females remain in dens for almost 8 weeks, emerging periodically to feed in the ninth to tenth week, and bring cubs out of the den shortly after. A litter size of two cubs is most common (Laurie & Seidensticker 1977; Phillips 1984; Joshi 1996). Although individual cubs have been observed, it is suspected that these are a result of early mortality of the second cub (Joshi 1996). Cubs stay with their mothers for 1.5–2.5 years, becoming independent just before the breeding season (Joshi et al. 1999). Thus, females breed at either 2- or 3-year intervals. The overall survival rate of cubs up to 2 years in Nepal was 69% (Joshi et al. 1999). Sloth bears exhibit mating seasonality. The mating season in sloth bears from mainland India spans May–July, and females are reported to give birth from November to January (Laurie & Seidensticker 1977; Gopal 1991; Joshi et al. 1999; Spady et al. 2007), usually after a period of delayed implantation (Puschmann et al. 1977). In Central India, young cubs have been observed on the back of females during February and March (Figure 2.2), which suggests that birth occurs in winter (Sharma and Dutta, personal observation). Phillips (1984) indicated that there was no conspicuous breeding season in Sri Lanka, although Ratnayeke (unpublished data from remote cameras and hunter surveys) concluded that most births in Sri Lanka matched the pattern reported for bears on the Indian mainland. An unusual characteristic of sloth bears is that females habitually carry the cubs (Norris 1969; Laurie & Seidensticker 1977; Joshi et al. 1999). Cubs ride on their mothers’ backs almost fulltime for the first 3 months (Figure 2.2), descending periodically to play and feed thereafter, but continue to be carried at least part of the time until they are 6–9 months old (Joshi et al. 1999). Cubs ride cross-wise or heads-forward, usually on the rump and lower back of the female. This behavior, which is shared by other myrmecophagous mammals, may have evolved independently in sloth bears to

24

Figure 2.2 A sloth bear cub riding on its mother’s back in Satpura National Park, India. (Photo by J. E. Swenson.)

guard against other large sympatric predators, such as tigers and leopards (Joshi et al. 1999). Tigers and leopards kill sloth bears occasionally (Kurt & Jayasuriya 1968; Joshi et al. 1999), and dholes and jackals could be a threat to cubs (Yoganand et al. 2013). Cubs can climb trees to feed on honey and fruits, but have never been observed to use trees to escape from predators (see also Chapter 9). In Chitwan National Park, Nepal, adult males and lone females are nocturnal, whereas females with cubs and subadults of both sexes are mostly diurnal, a behavior proposed to safeguard young from nocturnal predators and infanticidal conspecifics (Joshi et al. 1996). However, Yoganand (2005) found that activity patterns for females with dependent cubs did not differ from those of adult males and solitary adult females in the dry deciduous forest of Panna National Park, probably due to lower sloth bear density and the lack of concentrated rich sources of food.

Asiatic Black Bear The Asiatic black bear is a size-dimorphic and polygamous species (Yamamoto et al. 1998), with adult body mass typically ranging between 100 and 200 kg for males and 50 and 125 kg for females. Growth patterns differ by sex, with males growing until about 6–8 years of age and females until about 3–5 years (Nakamura et al. 2011). Males and females generally reach sexual maturity around 2–3 and 4 years old, respectively (Komatsu et al. 1994; Katayama et al. 1996), but females in nutritionally good condition can reproduce earlier. Although there are no data available, the age of first reproduction in the wild is probably delayed for males, because of strong male–male competition for females. Anecdotal reports suggest that females are reproductively active until their early twenties (Katayama et al. 1996; Nakamura et al. 2011; see also Chapter 23). In males, spermatogenesis starts during hibernation in March or April, and peaks in May and June to correspond with female receptivity (Komatsu et al. 1994). Although

Mating Strategies

detailed observations of mating behavior in the wild have not been reported, mating behavior and aggregations of adult bears have been observed in June and July. The mating season lasts from mid-June to early August for bears in captivity (Yamamoto et al. 1998). During the mating season, sexually mature males tend to be more active than females (Kozakai et al. 2013), and they search for receptive females and aggregate at locations where females are likely to be found (Oi et al. 2008). Males sometimes commit infanticide, which can induce mating opportunities with the victimized mother (Yamazaki 2017). The frequency of infanticide, its influence on cub mortality, and the survival rate of cubs-of-the-year is unknown (see also Chapter 10). Many males captured during this period carry wounds or injuries, indicating contest competition among males for females and/or sexual conflict. The length of the estrous period varies widely among females, ranging from 12 to 35 days, and estrous and anestrous periods tend to be repeated in an irregular pattern. Hence, multiple paternity litters can occur (Yamamoto et al. 2012). Females are believed to be induced ovulaters. Male reproductive behavior coincides with estrus. Younger males (4–5 years old) mount females less often that older males, and among adult males (>5 years old), the larger ones mate most frequently (Yamamoto et al. 1998). Females have delayed implantation, which occurs in late November–early December. Cubs are born from late January to early February. Newborn cubs weigh about 300 g. They grow fast, and accumulate around 2–3 kg of body mass over the approximately 90 days between birth and the end of hibernation (i.e. around 20–33 g/day). Based on observations of corpora lutea, corpora albicantia, and placental scars in killed bears, as well as the number of cubs observed in the wild, the mean number of ovulations per reproductive cycle is estimated to be 1.9, the mean number of implantations 2.0, and mean litter size 1.9 (Katayama et al. 1996). In contrast, the mean litter size in captive animals is reported to be 1.4 (Iibuchi et al. 2009). Although no observations of the interlitter interval in wild individuals currently exist, it is thought to be 2–3 years. Mothers and cubs-of-the-year remain in the vicinity of their place of hibernation until mid-May, and continue to remain together in the following months. The exact lactation period is unknown; however, lactation has been observed in adult females captured as late as October, meaning that it is possible that cubs are not weaned until autumn. Cubs-of-theyear are thought to overwinter with their mothers during the next winter and become independent during the breeding season of the following year, but the precise timing of the separation is unknown (see also Chapter 10).

American Black Bear The mating activities of American black bears (Ursus americanus) start in May, peak in June, and extend until July (Jonkel & Cowan 1971). Males usually are 10–20% longer and 20–70% heavier than females (Larivière 2001; Kovach & Powell 2003).

During a single mating season, males mate with multiple females and females mate with multiple males (Jonkel & Cowan 1971). Therefore, the American black bear is considered a seasonal polygamous breeder (Kovach & Powell 2003; see also Chapter 11). Male testicular activity starts during the late denning period in February or March. The peak of testicular activity occurs at the onset of the mating season and decreases slowly until October, after which the testicular activity ceases (Garshelis & Hellgren 1994). Males reach sexual maturity when around 3 years old, but they rarely successfully reproduce at that age because of competition with older and heavier males. Females’ age at first reproduction varies from 2 to 8 years, but most females produce their first litter at 5 or 6 years of age (Jonkel & Cowan 1971). Females are polyestrous with interestrous intervals of about 10 days, and multiple paternity litters are common (Himelright et al. 2014; Ombrello et al. 2016). Induced ovulation has been suggested (Boone et al. 1998, 2004). During receptive periods, females are usually followed by multiple males, but larger and heavier males tend to guard females during the peak of their estrus (Kovach & Powell 2003). Smaller and lighter males gain access to females mostly outside of the females’ receptive periods, or can obtain mating opportunities when multiple females experience estrus simultaneously (Kovach & Powell 2003; Costello et al. 2009). Consequently, paternity is often skewed toward a few dominant males in an area (Moore et al. 2015). Females have delayed implantation, which occurs in late November or early December (Tsubota et al. 1998). Once implanted, blastocyst(s) develop(s) for 60–70 days and parturition occurs in January or early February (Larivière 2001). Females produce litters of 1–4 cubs that weigh about 400 g each, and older and heavier females usually produce larger litters (Kolenosky 1990; Larivière 2001). Parental care is provided by the female only, which has lactational anestrus during the first mating season following cub birth (Larivière 2001). Lactation lasts for about 7 months, but cubs stay with their mother their entire first year. Cub survival varies from 25% to 50% and most cubs die from natural causes, such as disease, den flooding, infanticide, or predation by brown bears, wolves, coyotes, and bobcats (LeCount 1987; Kolenosky 1990; Larivière 2001). Family breakup occurs before or during the mating season, when offspring are approximately 16 months old (Jonkel & Cowan 1971). Depending on food availability and female body condition, litter intervals range from 1 to 4 years, but females usually breed every second year, as they can reproduce during the mating season following family breakup (Kolenosky 1990; Larivière 2001).

Brown Bear The brown bear has a polygamous mating system in the broadest sense of the term, as both sexes can associate and

25

Systematics, Ecology, and Behavior

mate with several partners during the mating season. Mating strategies, however, can vary across populations (Steyaert et al. 2012). It is a size-dimorphic species, in which adult males can be over twice the size of adult females. Such dimorphism is characteristic for species with strong sexual selection (Steyaert et al. 2012). Brown bears are seasonal breeders with a mating season that lasts from spring to early summer. The exact timing of the breeding season, however, varies across populations, as it correlates with photoperiod (Spady et al. 2007), and reproductive behavior during the autumn has occasionally been observed (Vaisefeld & Chestin 1993; see also Chapters 12 and 13). Brown bears, particularly males, have large home ranges that can extend over several 1000 km2 and overlap the ranges of many other males and females (McLoughlin et al. 2000). During the mating season, both sexes roam-to-mate, but especially males tend to increase their movements for doing so (Dahle & Swenson 2003; Krofel et al. 2010). Reproductive associations are usually comprised of one male and one female, but associations with several males and one female, or vice versa, are also common (Stenhouse et al. 2005). When traveling in associations, males monitor the females’ estrous status, which is probably signaled by pheromones (Craighead et al. 1995). Associations can last for just a few hours or up to several weeks. Single males sometimes sequester individual females at remote sites (e.g. mountain ridges) and over prolonged periods of time, which may be a strategy to assure paternity for the sequestering male (Hamer & Herrero 1990; Edwards & Derocher 2015). Multiannual mating areas as small as just few hectares may develop in small or low-density populations (e.g. Cantabrian Mountains, Spain; see also Chapter 19). Such areas can facilitate individuals to find each other and reproduce, and counteract the challenge of finding mates at low population densities (Fernández-Gil et al. 2006; Fernández-Gil 2013). In populations where bears aggregate around clustered food resources during the mating season (e.g. garbage dumps), social hierarchies can develop, where the most dominant male can monopolize reproductive females and prevent subdominant males from mating (Craighead et al. 1995). Strong competition for females exists among males. Prior to copulation, both scramble and contest competition occur, in which the latter can inflict severe injuries or cause death, especially in younger or smaller individuals (Craighead et al. 1995). Body size and condition, age, and aggression typically determine the dominance status of males (Zedrosser et al. 2007). Competition after copulation may occur as sperm competition within the reproductive tract of the female brown bears, although no conclusive evidence for this exists (Bellemain et al. 2006b). Competition can even extend after females give birth, as sexually selected infanticide (SSI) (Bellemain et al. 2006a; Davoli et al. 2018), a reproductive strategy in which males kill dependent conspecific offspring for obtaining mating opportunities (Hrdy 1979). Females control partner choice and mating to some extent and sometimes initiate mating. In populations with SSI, they face a dilemma between

26

selecting for the best-quality males or mating with all nearby males to reduce the risk for infanticide (Bellemain et al. 2006b), as promiscuity may confuse paternity and is considered as a counterstrategy to infanticide (Ebensperger 1998). Larger, older, and more heterozygous males that live in the vicinity of a female typically sire her offspring (Bellemain et al. 2006b; Zedrosser et al. 2007), and many males do not reproduce during a given mating season (Steyaert et al. 2012). Reproductive parameters show large variation among populations. Those parameters are typically related to habitat quality, which varies tremendously across the brown bears’ geographical distribution, from rich temperate ecosystems with access to abundant spawning salmon to harsh desert ecosystems (Steyaert et al. 2012; see also Chapters 12 and 13). Males reach sexual maturity around the age of 5–6 years, but occasionally reproduce at a younger age (Zedrosser et al. 2007). Testicular activity has a circannual rhythm, with testes size and function being greatest during the mating season (Spady et al. 2007). Female age of primiparity generally varies between about 5 and 10 years of age, although some females give birth at 3 years of age (Steyaert et al. 2012). Their prime reproductive age is between 9 and 20 years of age, and post reproductive survival is limited (Schwartz et al. 2003). Females are polyestrous and multiple paternity litters are common (Bellemain et al. 2006b). Females have lactational anestrus, and can (re)enter estrus only a few days after litter loss during the mating season (Steyaert et al. 2014). Induced ovulation has been suggested. Delayed implantation occurs in November–December and parturition occurs after 6–8 weeks of active gestation, around January– February (Spady et al. 2007; Friebe et al. 2014). Females usually give birth to 1–3 altricial cubs, but litter sizes up to 6 have been reported. Mean interlitter intervals typically range between 2.4 and 5.7 years and maternal care lasts for about 1.4–3.5 years in most populations (Figure 2.3), but can be as long as 4.5 years (Steyaert et al. 2012).

Figure 2.3 A female brown bear with her yearling at the beginning of the autumn. (Photo by V. Penteriani.)

Mating Strategies

Polar Bear The mating season of the polar bear may vary geographically, but ranges between March and June (Smith & Aars 2015). The mating system is best described as serial female defense polygyny (Ramsay & Stirling 1986). A form of a polygynous mating system is supported by marked sexual dimorphism, with males typically 2–3 times heavier than females (Derocher et al. 2010). Due to the prolonged mother–offspring bond of about 2.5 years, fewer females are available to breed in any given year than there are males (2–3 males per female) (Molnár et al. 2008) (see also Chapter 14). The male-biased operational sex ratio results in intense intrasexual conflicts for access to estrous females and these are noted in tooth breakage (mainly canines), cuts, and scars (Ramsay & Stirling 1986; Derocher et al. 2010). Injuries can be severe and can include, for example, broken bones and loss of vision (Derocher et al. 2010). Polar bears live at low densities and have large overlapping home ranges (usually >100,000 km2; Mauritzen et al. 2001). This low density, in addition to low predictability of mates and dynamic sea ice conditions, results in males performing longdistance searches for available females (Stirling et al. 2016). Males track females over long distances, presumably to assess their breeding status based on scents in the tracks (Owen et al. 2015). Once a male locates an estrous female, he may try to herd her toward a remote area, or keep her in a small area to reduce the likelihood of detection by other males (Wiig et al. 1992; Stirling et al. 2016). Most observations of mating occur on the sea ice (Figure 2.4), but some pairs move onto land. Males in mated pairs tend to be slightly older than unpaired males (Derocher et al. 2010), but the importance of age may vary by population (Zeyl et al. 2009). Estrous females are usually accompanied by a single male, but observations of 2–5 males near an estrous female are common. Successful access to a female is thought to occur largely through contest competition, although some pairings may result from scramble competition (Derocher et al. 2010). Female mate

choice is poorly understood in polar bears, but the wide range in body mass of males with females (1:1 to 3:1) suggests that it may be limited. Male–female associations normally last about 2 weeks, but can range up to 18 days (Stirling et al. 2016). Males can produce sperm at 2 years of age (Rosing-Asvid et al. 2002), but rarely sire offspring until the age of 6 years, after which a steep increase in mating success can occur. Spermatogenesis starts in February and testes size is largest during the breeding season. Male mating success peaks at around 14 years of age, after which a steep decline occurs (Richardson 2014). Lifetime mating success in male polar bears is heavily skewed, with only few males siring many offspring (up to at least 9 successful matings; Richardson 2014). Females typically mate first when 4–6 years old and produce their first litter the following year (Ramsay & Stirling 1988). An increase in reproduction occurs until about 14 years of age, followed by a slow decline, and they can reproduce until their late 20s (Derocher & Stirling 1994). Polar bears are thought to be induced ovulators, and undergo delayed implantation during autumn. Females may switch partners while in estrus, which can result in multiple paternity litters (Zeyl et al. 2009). Inbreeding in polar bears appears to be low (Zeyl et al. 2009). Parturition occurs between November and January and the cubs are altricial (Amstrup & DeMaster 2003). Litter size depends on maternal age and body condition and varies from one to four, with singletons or twins being most common (Ramsay & Stirling 1988). Adoption sometimes occurs (Derocher & Wiig 1999b). Estimates of cub survival range from below 20% to over 95% between years (Bromaghin et al. 2015) and are influenced mainly by environmental conditions and maternal factors (Derocher & Stirling 1996). Although the typical reproductive interval for females is three years, intervals of one or two years are common if a litter is lost (Ramsay & Stirling 1988). Infanticide occurs in polar bears and appears associated with nutritional gain, as it usually occurs outside the mating season (Derocher & Wiig 1999a; Stone & Derocher 2007). Therefore, there is limited evidence for sexually selected infanticide as a reproductive tactic. As with other ursids, there is no paternal investment in rearing of young.

Chemical Communication

Figure 2.4 A mating pair of polar bears during spring in the Beaufort Sea, Canada. (Photo by A. E. Derocher.)

Chemical communication involves sending and receiving signals and cues via organic chemicals. Chemicals exit the body through glandular secretions and excretions and are either passively released or actively directed toward environmental objects, forming a scent mark (Adams 1980). Scent marking is particularly prevalent in mammals and functions broadly in territory defense, group cohesion (for social species), recognition of individuals and kin, and assessment of mates or competitors (Muller-Schwarze 2006). Bears have evolved a variety of physical traits and behavioral strategies to communicate efficiently using organic compounds. Glandular secretions from modified sweat and

27

Systematics, Ecology, and Behavior

the females’ energetic investment in gamete production is not wasted and reduces the risk to males by reducing the necessity for physical contests. Large gaps in our knowledge of the function of chemical signaling and its role in the mating systems of bears remain, particularly for those species inhabiting tropical regions. The apparent similarities in chemical signaling and marking behavior among several bears suggest that the chemical signaling in sexual behavior is an ancestral trait of the Ursidae.

Reproductive Physiology Figure 2.5 Scent marking in bears is biased toward males. Here, however, an Alaskan female brown bear with her three cubs-of-the-year investigated and scent marked a rubbing tree. After the mother rubbed the tree, the cubs imitated her behavior, suggesting that female bears may teach their cubs to scent-mark. (Photo by E. Filipczykova.)

sebaceous glands appear to be the primary sources of scent used for marking, along with urine (Sergiel et al. 2017). Bears are selective in where, how, and when they place scent marks within their environment (Clapham et al. 2014). For those bear species studied, all appear to increase marking behavior around the mating season. Observations indicate that adult males scent mark more than other age–sex classes (Tattoni et al. 2015; Figure 2.5), although this varies in giant pandas. Due to the general lack of territoriality in bears and their solitary living, chemical communication appears to be focused primarily on sexual behavior. Giant pandas have been the focus of most of the research conducted on identifying the chemical signals and cues contained within bear odor, which include: sex, age, estrous state, kinship, and individuality (Swaisgood et al. 2002; Liu et al. 2008). In addition, compounds coding for sex have been identified in brown bears (Rosell et al. 2011) and behavioral discrimination of sex and estrous state has been shown in polar bears (Owen et al. 2015). The role of chemical signaling in the mating system of bears appears to function to facilitate reproductive advertisement and male intrasexual competition. Female bears seem to rely on chemical cues to advertise reproductive condition, and males likely select and locate mates based on this knowledge (Clapham et al. 2012; Owen et al. 2015). Behavioral studies suggest that male bears use scent marking to signal competitive ability to other males for access to females, which may involve year-round signaling of dominance (Nie et al. 2012a). By signaling competitive ability, male bears may be able to reduce the risk of injury from physical aggression by influencing competitors into subordination, as seen in other mammals. Population density, environmental factors, and reproductive biology all likely play a role in the varying strategies that bears use to communicate chemically (Owen et al. 2015; Lamb et al. 2017). In evolutionary terms, chemical signaling ensures that

28

Bears possess a number of physiologic adaptations of the reproductive system that allow them to increase reproductive success and maximize genetic diversity. The ovary is largely responsible for most of these physiologic adaptations in bears, including delayed implantation, split parturition, and pseudopregnancy. The bear ovary is covered by a complete bursa (Erickson et al. 1964), which likely reduces egg loss and increases chances of successful fertilization. Bear ovaries frequently release multiple eggs at once and, except in the giant panda, also can ovulate sequentially during the mating season. Unlike most other polyestrous mammals, their ovaries continue to ovulate when the female is pregnant (during early embryonic diapause) and/or has already ovulated that mating season (Himelright et al. 2014). This helps to explain the relatively high proportion (15–20%) of multipaternity litters observed in high-density, free-ranging bear populations (Bellemain et al. 2006b; Ombrello et al. 2016; Shimozuru et al. 2019). Multipaternity, in turn, likely serves to increase genetic diversity, mitigate risks of inbreeding depression, and may provide a paternity confusion strategy to cope with infanticidal males (Steyaert et al. 2012). Interestingly, the corpora lutea (CL) of ovaries in all eight bear species remain relatively inactive until 0.5–6.5 months after ovulation, depending on the species, at which point all CL in a female’s ovaries synchronously reactivate, regardless of when individual ovulations occurred that season (Spady et al. 2007). This CL dormancy feature is unique to carnivores with delayed implantation (bears, wolverines, badgers, etc.). In bears specifically, the length of delay until implantation is highly variable within and between individuals. Among seasonally breeding bear species, however, the timing of implantation within the year is consistent and associated with latitude (i.e. late October/early November in brown and American black bears in the Temperate Zone; Spady et al. 2007). At the species level, the duration of delay is similarly long between five of the seven species examined to date (10–27 weeks; Spady et al. 2007). In contrast, the delay in giant pandas is slightly shorter on average (7–15.5 weeks; Zhang et al. 2009). Whether sun bears exhibit delayed implantation remains unclear. Although past studies concluded that sun bears lack delayed implantation, data indicate that a period of at least 2–4 weeks of CL dormancy after ovulation occurs (see Schwarzenberger

Mating Strategies

et al. 2004; Frederick et al. 2012). Because implantation is preceded by active luteal function, sun bear embryos may exhibit a short period of diapause lasting at least 2 weeks. Both long and short delayed implantation have been hypothesized as mechanisms for enhancing reproductive fitness via increased multipaternity and cryptic female choice (Birkhead & Møller 1993). Hence, even the aseasonal sun bears would benefit from this. The fertilized bear egg grows and develops normally after conception until it reaches the preimplantation blastocyst stage, at which point growth drastically slows and development is arrested (Himelright et al. 2014). The end of delay begins with CL reactivation, resulting in increased progestins, which initiates the uterine secretory phase. Only then do the diapaused blastocysts implant and resume development. Because birth occurs just 6–9 weeks after implantation, depending on the species, cubs are altricial at birth, with giant panda cubs the least developed (Mead 1993; Tsubota & Kanagawa 1993). Other polyestrous mammals do not have this delay in CL activity; therefore, once a mating results in pregnancy, high progestin levels produced by the active CL block any subsequent follicular development and estruses that same season. Accordingly, sequential ovulations during the mating season of most polyestrous species only occur if the previous estrus did not result in a pregnancy. Hence, without some physiological means to do so, polyestrous species cannot become pregnant following successful fertilization during a previous ovulation. This is not true for bears, however, thus it is theoretically possible for them to become pregnant while they are already pregnant (superfetation). Although superfetation has not yet been proven in bears, it has been documented in related Carnivora with delayed implantation. Superfetation is further supported by the observations that ovulations can occur weeks apart in polyestrous bears (Himelright et al. 2014), and that all CL present reactivate in synchrony after the mating season (Spady et al. 2007). In addition, within 1–2 weeks after CL reactivation, the uterus has achieved the secretory phase, and all diapaused embryos implant and develop in synchrony (Tsubota & Kanagawa 1993; Yamane et al. 2009). In theory, these embryos may have been conceived by different sires at different times in the season, but because they all implant at the same time, they would be born into the same litter and of the same relative developmental age. Giant pandas and sun bears are spontaneous ovulators (Durrant et al. 2002; Frederick et al. 2013), but in the other six bear species, ovulation is presumed to be induced by physical stimulation of vaginal nerves during coitus (Boone et al. 1998). However, subsequent studies (e.g. Okano et al. 2006) support the conclusion that spontaneous ovulation, or ovulation induced by triggers other than coitus (such as other physical contact, or perhaps olfactory or visual cues), is the primary mode of ovulation in bears. All bear species exhibit pseudopregnancy, which is a consequence of ovulation without subsequent fertilization of ova. Pseudopregnant females share many behavioral and physiological characteristics with

pregnant females, such as nest-building behavior, uterine hypertrophy, and mammary development, all in response to prolonged exposure to elevated progestins and other hormones, but without the embryo (Concannon 2009).

Sexually Selected Infanticide (SSI) SSI is a male reproductive strategy in which an adult male kills (or induces abortion) of unrelated offspring to shorten the time to when he can impregnate the mother, and thus increases his fitness (Hrdy 1979). SSI is common across many taxa. In terrestrial mammals, it has been most clearly studied in rodents, primates, and carnivores. The three requirements for SSI to evolve are simple: (1) the interbirth interval of the female should be shortened if dependent offspring are killed; (2) the probability of a male killing his own offspring should be very low; and (3), the male, or his close relative, should have a high probability of fathering the mother’s next offspring. SSI would benefit the fitness of the male, but is very costly to the fitness of the mother, and thus female counterstrategies would also be expected to have evolved (Ebensperger 1998). These counterstrategies primarily include: (1) female aggression and coalitions to defend offspring; (2) mothers avoiding males that are too large to repel; (3) promiscuity to confuse paternity so males tolerate young; and (4) assistance in defending the territory of a family group. The entire SSI process is extremely difficult to document in bears, as they are generally solitary, have large and overlapping home ranges, and are often difficult to observe. Among bears, SSI has mostly been studied in the brown bear, in which it can induce >30% cub loss per year (Gosselin et al. 2017; Figure 2.6). In general, bears clearly meet the first requirement for SSI, as they have extended interbirth intervals, during which mothers do not enter estrus (lactational anestrus) or mate. Killing cubs, or even yearlings, would return the mother to a breeding condition earlier. In Scandinavian brown bears, for example, females typically enter estrus just a few days after losing an entire litter during the mating season (Steyaert et al. 2014). The second and third requirements are usually not as clear, although some genetic evidence for both requirements exists in brown bears (Bellemain et al. 2006a; Davoli et al. 2018). Unless a male has never mated before, immigrates into an area, or can recognize the females he mated with the previous year, there may be a nontrivial probability that an infanticidal male could kill his own offspring, particularly if there were few other males in the population. However, males can undoubtedly recognize females, as individual recognition is widespread in a great variety of animals (Tibbetts & Dale 2007) and bears, like many mammals, spend considerable effort physically and chemically advertising their presence (Lamb et al. 2017). Furthermore, the female counterstrategy of confusing paternity by mating with several males would only work if males can recognize individual females they have mated with (Bellemain et al. 2006b; Fernández-Gil 2013). The second requirement for SSI is thus likely met in bears, with

29

Systematics, Ecology, and Behavior

Figure 2.6 Male brown bears sometimes consume their infanticide victims, as the cub claw and hair in the feces depicted evidences. Here, a GPS-marked adult male killed and ate a cub-of-the-year of a GPS-marked female during the mating season of 2010 in south-central Sweden. (Photo by S.M.J.G. Steyaert.)

males not killing the cubs of mothers that they mated with the previous year. The third SSI requirement, where the infanticidal male has a high probability of fathering the victimized mother’s next offspring, is least clear. Not being territorial, male bears do not have exclusive breeding rights with any female. However, as females can rapidly enter estrus after losing an entire litter, consorting or guarding such females until estrus could give males (even if not dominant) an opportunity to mate and to father the female’s next offspring. Because bear density varies by over two orders of magnitude, there are likely areas with a low enough density of adult male bears that even smaller infanticidal males can successfully consort and mate with the female. The third requirement of the SSI hypothesis is thus likely met in some species or populations, particularly for large, dominant males.

Human Impacts on the Mating System of Bears Bears typically roam over large areas during the mating season (Dahle & Swenson 2003), which can increase the probability of coming into contact with humans or human-related infrastructures, compared to other periods of the year (Penteriani et al. 2018). For instance, road mortality of brown bears in Europe is higher during the mating season (Kaczensky et al. 2003), when bears likely cross roads more often than when they have smaller home ranges. The despotic nature of male bears is most pronounced during the mating season. More vulnerable bears, such as subadults and family groups, may avoid male aggression during the mating season by using areas relatively close to humans, which also make them more prone to become involved in human–bear interactions and conflict (Elfström et al. 2014; Steyaert et al. 2016). Some human activities focus on sensitive places where bears congregate during the mating season (Fernández-Gil et al. 2006) and where females with cubs try to stay away from other bears. That is the case of ecotourism focused on the observation of bears

30

(Penteriani et al. 2017). Bears may leave such important areas or reduce activity during daytime because of human disturbance, which in turn can have negative effects on mating, cub rearing, and, ultimately, fitness (Fortin et al. 2016; Penteriani et al. 2017). It also remains unclear how male–female associations during the mating season occur, to what extent human activities and infrastructures limit roam-to-mate behavior, and how that influences reproductive success. A better understanding of where male and female bears interact and how they move during the mating season would allow the quantification of the influence of anthropogenic disturbances on the mating system of various bear species. Of importance from a conservation and management point of view is that trophy-hunting of adult males can cause rapid male turnover and home range rearrangement, which may artificially increase SSI and can limit population growth. This mechanism is controversial, as it cannot be generalized among bear species and populations. For example, strong evidence for this mechanism exists in a population in south-central Sweden (Swenson et al. 1997; Gosselin et al. 2015, 2017; Leclerc et al. 2017). Studies in North America, however, have found that removing adult males had no discernible effect on the survival of cubs of mothers that lived within the dead males’ home range (McLellan 2005, 2015; Schwartz et al. 2006) and cub survival is lowest in populations where adult male survival is highest (McLellan, unpublished data). Furthermore, an experimental removal of male American black bears did not show a change in cub survival (Czetwertynski et al. 2007). Undoubtedly, demographic parameters (e.g. population density, operational sex ratio, size dimorphism), as well as ecological context (e.g. availability and clustering of food resources), determine the prevalence of SSI in bear species and populations and how male removal impacts cub survival. How and to what extent, however, remains unclear. Depletion of male bears, for example due to overharvest, may also affect their mating system by reducing the encounter rates of estrous females, as has been documented in polar bears (Molnár et al. 2008). As for many other wildlife species, habitat destruction and fragmentation can have far-reaching effects on behavior and life history in bears. In giant pandas, for example, females rely on dens for birth and early cub development. Due to the fragility of the cub, selection of dens can be very important to provide protection from predators and buffer the cub from inclement weather (Wei et al. 2019). Female pandas select dens in tree hollows and rock dens based on characteristics that will provide a warm, dry, stable environment, such as the size of the opening. Given widespread logging throughout much of the panda’s range, the quantity and quality of tree hollows for rearing cubs has been diminished, and may explain why pandas favor old-growth to secondary-growth forests (Zhang et al. 2011). Thus, a variety of human activities can have negative effects on bears during the mating season, particularly in humandominated, fragmented landscapes. In addition, climate change (Derocher et al. 2004; Hertel et al. 2018) and other

Mating Strategies

human-induced disturbances of ecosystems (e.g. pollution) can affect mating systems and strategies. To what extent human impacts affect the mating systems and strategies of bears remains largely unknown. Improving our knowledge

References Adams, M. (1980). Odour-producing organs of mammals. Symposia of the Zoological Society of London 45: 57–86. Amstrup, S. C. & DeMaster, D. P. (2003). Polar bear, Ursus maritimus. Wild Mammals of North America: Biology, Management, and Conservation 2: 587–610.

about how humans impact mating systems and strategies of bears and wildlife in general is paramount for effective management and conservation, especially for small and endangered populations.

patterns? International Bear News 27: 57–58.

female polar bears. Journal of Zoology 234: 527–536.

Clapham, M., Nevin, O. T., Ramsey, A. D. & Rosell, F. (2012). A hypotheticodeductive approach to assessing the social function of chemical signalling in a nonterritorial solitary carnivore. PLoS ONE 7: e35404.

Derocher, A. E. & Stirling, I. (1996). Aspects of survival in juvenile polar bears. Canadian Journal of Zoology 74: 1246–1252. Derocher, A. E. & Wiig, Ø. (1999a). Infanticide and cannibalism of juvenile polar bears (Ursus maritimus) in Svalbard. Arctic 52: 307–310.

Andersson, M. (1994). Sexual selection. Princeton, NJ: Princeton University Press.

Clapham, M., Nevin, O. T., Ramsey, A. D. & Rosell, F. (2014). Scent-marking investment and motor patterns are affected by the age and sex of wild brown bears. Animal Behaviour 94: 107–116.

Appleton, R. D., Van Horn, R. C., Noyce, K. V., et al. (2018). Phenotypic plasticity in the timing of reproduction in Andean bears. Journal of Zoology 305: 196–202.

Concannon, P. (2009). Endocrinologic control of normal canine ovarian function. Reproduction in Domestic Animals 44: 3–15.

Bellemain, E., Swenson, J. E. & Taberlet, P. (2006a). Mating strategies in relation to sexually selected infanticide in a nonsocial carnivore: the brown bear. Ethology 112: 238–246.

Derocher, A. E., Lunn, N. J. & Stirling, I. (2004). Polar bears in a warming climate. Integrative and Comparative Biology 44: 163–176.

Connor, T., Hull, V. & Liu, J. (2016). Telemetry research on elusive wildlife: a synthesis of studies on giant pandas. Integrative Zoology 11: 295–307.

Derocher, A. E., Andersen, M., Wiig, Ø. & Aars J. (2010). Sexual dimorphism and the mating ecology of polar bears (Ursus maritimus) at Svalbard. Behavioral Ecology and Sociobiology 64: 939–946.

Bellemain, E., Zedrosser, A., Manel, S., et al. (2006b). The dilemma of female mate selection in the brown bear, a species with sexually selected infanticide. Proceedings of the Royal Society B 273: 283–291.

Costello, C. M., Creel, S. R., Kalinowski, S. T., Vu, N. V. & Quigley H. B. (2009). Determinants of male reproductive success in American black bears. Behavioral Ecology and Sociobiology 64: 125.

Birkhead, T. R. & Møller, A. P. (1993). Sexual selection and the temporal separation of reproductive events: sperm storage data from reptiles, birds and mammals. Biological Journal of the Linnean Society 50(4): 295–311.

Craighead, J. J., Sumner, J. S. & Mitchell, J. A. (1995). The grizzly bears of Yellowstone: Their ecology in the Yellowstone Ecosystem, 1959–1992. Washington, DC: Island Press.

Boone, W. R., Catlin, J. C., Casey, K. J., et al. (1998). Bears as induced ovulators: a preliminary study. Ursus 10: 503–505. Boone, W. R., Keck, B. B., Catlin, J. C., et al. (2004). Evidence that bears are induced ovulators. Theriogenology 61: 1163–1169. Bromaghin, J. F., McDonald, T. L., Stirling, I., et al. (2015). Polar bear population dynamics in the southern Beaufort Sea during a period of sea ice decline. Ecological Applications 25: 634–651. Castellanos, A. & Vasquez, D. (2017). Kinship relations in a multi-generational Andean bear (Tremarctos ornatus) family in North Ecuador. International Bear News 26: 24–25. Castellanos, A., Ascanta, M. & Jackson, D. (2018). Does Rebecca, a seasoned Andean bear mother, show seasonal birthing

Crudge, B., Lees, C., Hunt, M., et al. (2019). Sun bears: Global status review and conservation action plan, 2019–2028. IUCN SSC Bear Specialist Group. Czetwertynski, S. M., Boyce, M. S. & Schmiegelow, F. K. (2007). Effects of hunting on demographic parameters of American black bears. Ursus 18: 1–18. Dahle, B. & Swenson, J. E. (2003). Seasonal range size in relation to reproductive strategies in brown bears Ursus arctos. Journal of Animal Ecology 72: 660–667. Davoli, F., Cozzo, M., Angeli, F., Groff, C. & Randi, E. (2018). Infanticide in brown bear: a case-study in the Italian Alps – genetic identification of perpetrator and implications in small populations. Nature Conservation 25: 55–75. Derocher, A. E. & Stirling, I. (1994). Agespecific reproductive performance of

Derocher, A. E. & Wiig, Ø. (1999b). Observation of adoption in polar bears (Ursus maritimus). Arctic 52: 413–415.

Durrant, B., Czekala, N., Olson, M., et al. (2002). Papanicolaou staining of exfoliated vaginal epithelial cells facilitates the prediction of ovulation in the giant panda. Theriogenology 57: 1855–1864. Ebensperger, L. A. (1998). Strategies and counterstrategies to infanticide in mammals. Biological Reviews 73: 321–346. Edwards, M. A. & Derocher, A. E. (2015). Mating-related behaviour of grizzly bears inhabiting marginal habitat at the periphery of their North American range. Behavioural Processes 111: 75–83. Elfström, M., Zedrosser, A., Jerina, K., et al. (2014). Does despotic behavior or food search explain the occurrence of problem bears in Europe? Journal of Wildlife Management 78(5): 881–893. Emlen, S. T. & Oring, L. W. (1977). Ecology, sexual selection and the evolution of mating systems. Science 197: 215–223. Enciso, M. A. & Guimarães, M. A. B. V. (2013). Knowing the reproductive endocrinology in the female Andean bear through non-invasive methods. International Bear News 22: 33–34. Erickson, A. W., Nellor, J. & Petrides, G. A. (1964). The black bear in Michigan.

31

Systematics, Ecology, and Behavior

Michigan State University Agriculture Experiment Station East Lansing Research Bulletin 4: 1–102. Fernández-Gil, A. (2013). Behavior and conservation of large carnivores in human-dominated landscapes. Brown bears and wolves in the Cantabrian Mountains. PhD thesis, University of Oviedo, Spain. Fernández-Gil, A., Naves, J. & Delibes, M. (2006). Courtship of brown bears Ursus arctos in northern Spain: phenology, weather, habitat and durable mating areas. Wildlife Biology 12: 367–373. Filipczyková, E., Heitkönig, I. M. A., Castellanos, A., Hantson, W. & Steyaert, S. M. J. G. (2017). Marking behavior of Andean bears in an Ecuadorian cloud forest: a pilot study. Ursus 27(2): 122–128. Fortin, J. K., Rode, K. D., Hilderbrand, G. V., et al. (2016). Impacts of human recreation on brown bears (Ursus arctos): a review and new management tool. PLoS ONE 11: e0141983. Frederick, C., Kyes, R., Hunt, K., et al. (2010). Methods of estrus detection and correlates of the reproductive cycle in the sun bear (Helarctos malayanus). Theriogenology 74: 1121–1135. Frederick, C., Hunt, K. E., Kyes, R., Collins, D. & Wasser, S. K. (2012). Reproductive timing and aseasonality in the sun bear (Helarctos malayanus). Journal of Mammalogy 93: 522–531. Frederick, C., Hunt, K., Kyes, R., et al. (2013). Social influences on the estrous cycle of the captive sun bear (Helarctos Malayanus). Zoo Biology 32: 581–591. Fredriksson, G. M. & Wich, S. A. (2006). Frugivory in sun bears (Helarctos malayanus) is linked to El Niño-related fluctuations in fruiting phenology, East Kalimantan, Indonesia. Biological Journal of the Linnean Society 89: 489–508. Fredriksson, G. M., Danielsen, L. S. & Swenson, J. E. (2007). Impacts of El Niño related drought and forest fires on sun bear fruit resources in lowland dipterocarp forest of East Borneo. Biodiversity and Conservation 16: 1823–1838. Friebe, A., Evans, A. L., Arnemo, J. M., et al. (2014). Factors affecting date of implantation, parturition, and den entry estimated from activity and body temperature in free-ranging brown bears. PLoS ONE 9(7): e101410. Garcia-Rangel, S. (2012). Andean bear Tremarctos ornatus natural history and

32

conservation. Mammal Review 42: 85–119. Garshelis, D. L. (2009). Family Ursidae (bears). In: Wilson, D. E. & Mittermeier, R.A. (Eds.), Handbook of the mammals of the world, Vol. 1 – Carnivores (pp. 448–497). Barcelona: Lynx Edicions. Garshelis, D. L. & Hellgren, E. C. (1994). Variation in reproductive biology of male black bears. Journal of Mammalogy 75: 175–188. Garshelis, D. L., Joshi, A. R., Smith, J. L. & Rice C. G. (1999). Sloth bear conservation action plan in bears: Status survey and conservation action plan (pp. 225–240). Gland: International Union for the Conservation of Nature and Natural Resources. Gittleman, J. L. (1994). Are the pandas successful specialists or evolutionary failures? BioScience 44: 456–464. Gopal, R. (1991). Ethological observations on the sloth bear (Melursus ursinus). Indian Forester 117: 915–920. Gosselin, J., Zedrosser, A., Swenson, J. E. & Pelletier, F. (2015). The relative importance of direct and indirect effects of hunting mortality on the population dynamics of brown bears. Proceedings of the Royal Society B: Biological Sciences 282: 20141840. Gosselin, J., Leclerc, M., Zedrosser, A., et al. (2017). Hunting promotes sexual conflict in brown bears. Journal of Animal Ecology 86: 35–42. Hamer, D. & Herrero, S. (1990). Courtship and use of mating areas by grizzly bears in the Front Ranges of Banff National Park, Alberta. Canadian Journal of Zoology 68: 2695–2697. Hertel, A. G., Bischof, R., Langval, O., et al. (2018). Berry production drives bottomup effects on body mass and reproductive success in an omnivore. Oikos 127(2): 197–207. Himelright, B. M., Moore, J. M., Gonzales, R. L., et al. (2014). Sequential ovulation and fertility of polyoestrus in American black bears (Ursus americanus). Conservation Physiology 2: cou051. Hrdy, S. B. (1979). Infanticide among animals: a review, classification, and examination of the implications for the reproductive strategies of females. Ethology and Sociobiology 1: 13–40. Hunter, L. & P. Barrett. (2019). Carnivores of the world. Second edition. Princeton, NJ: Princeton University Press.

Iibuchi, R., Nakano, N., Nakamura, T., et al. (2009). Change in body weight of mothers and neonates and in milk composition during denning period in captive Japanese black bears (Ursus thibetanus japonicus). Japanese Journal of Veterinary Research 57: 13–22. Jonkel, C. & Cowan, I. (1971). The black bear in the spruce-fir forest. Wildlife Monographs 27: 3–57. Joshi, A. R. (1996). The home range, feeding habits, and social organization of sloth bears (Melursus ursinus) in Royal Chitwan National Park, Nepal. PhD thesis, University of Minnesota, Minneapolis, USA. Joshi, A. R., Smith, J. L. D. & Garshelis, D. L. (1999). Sociobiology of the myrmecophagous sloth bear in Nepal. Canadian Journal of Zoology 77(11): 1690–1704. Kaczensky, P., Knauer, F., Krze, B., et al. (2003). The impact of high speed, high volume traffic axes on brown bears in Slovenia. Biological Conservation 111: 191–204. Katayama, A., Tsubota, T., Yamada, F., Kita, I. & Tiba, T. (1996). Reproductive evaluation of Japanese black bears (Selenarctos thiberanus japonicus) by observation of the ovary and uterus. Japanese Journal of Zoo Wildlife Medicine 1: 26–32. Kleiner, J. D., Van Horn R. C., Swenson J. E. & Steyaert, S. M. J. G. (2018). Rub-tree selection by Andean bears in the Peruvian dry forest. Ursus 29(1): 58–66. Kolenosky, G. B. (1990). Reproductive biology of black bears in east-central Ontario. Bears: Their Biology and Management 8: 385–392. Komatsu, T., Tsubota, T., Kishimoto, M., Amasaki, S. & Tiba, T. (1994). Puberty and stem cell for the initiation and resumption of spermatogenesis in the male Japanese black bear (Selenarctos thibetanus japonicus). Journal of Reproduction and Development 40: j65–j71. Kovach, A. I. & Powell, R. A. (2003). Effects of body size on male mating tactics and paternity in black bears, Ursus americanus. Canadian Journal of Zoology 81: 1257–1268. Kozakai, C., Yamazaki, K., Nemoto, Y., et al. (2013). Fluctuation of daily activity time budgets of Japanese black bears: relationship to sex, reproductive status, and hardmast availability. Journal of Mammalogy 94: 351–360.

Mating Strategies

Krofel, M., Filacorda, S. & Jerina, K. (2010). Mating-related movements of male brown bears on the periphery of an expanding population. Ursus 21: 23–30. Kurt, F. & Jayasuriya, A. (1968). Notes on a dead bear. Loris 11: 182–183. Lamb, C. T., Mowat, G., Gilbert, S. L., et al. (2017). Density-dependent signaling: an alternative hypothesis on the function of chemical signaling in a non-territorial solitary carnivore. PLoS ONE 12: e0184176. Larivière, S. (2001). Ursus americanus. Mammalian Species 647: 1–11. Laurie, A. & Seidensticker, J. (1977). Behavioural ecology of the sloth bear (Melursus ursinus). Journal of Zoology 182 (2): 187–204. Leclerc, M., Frank, S. C., Zedrosser, A., Swenson, J. E. & Pelletier, F. (2017). Hunting promotes spatial reorganization and sexually selected infanticide. Scientific Reports 7: 45222.

in home range overlap with habitat quality: a comparison among brown bear populations. Evolutionary Ecology 14: 39–60. Mead, R. A. (1993). Embryonic diapause in vertebrates. Journal of Experimental Zoology 266: 629–641. Molnár, P. K., Derocher, A. E., Lewis, M. A. & Taylor, M. K. (2008). Modelling the mating system of polar bears: a mechanistic approach to the Allee effect. Proceedings of the Royal Society B – Biological Sciences 275: 217–226. Moore, J. A., Xu, R., Frank, K., Draheim, H. & Scribner, K. T. (2015). Social network analysis of mating patterns in American black bears (Ursus americanus). Molecular Ecology 24: 4010–4022. Muller-Schwarze, D. (2006). Chemical ecology of vertebrates. Cambridge: Cambridge University Press.

bears (Ursus thibetanus japonicus) based on serum progesterone profiles. Journal of Veterinary Medical Science 68: 1133–1137. Ombrello, T. A., Chinnici, N. L. & Huffman, J.E. (2016). Multiple paternities in American black bears from New Jersey. Journal of the Pennsylvania Academy of Science 90: 21–24. Owen, M. A., Swaisgood, R. R., Slocomb, C., et al. (2015). An experimental investigation of chemical communication in the polar bear. Journal of Zoology 295: 36–43. Pan, W. (2014). A chance for lasting survival: Ecology and behavior of wild giant pandas. Washington, DC: Smithsonian Institution Scholarly Press. Penteriani, V., López-Bao, J. V., Bettega, C., et al. (2017). Consequences of brown bear viewing tourism: a review. Biological Conservation 206: 169–180.

LeCount, A. L. (1987). Causes of black bear cub mortality. Bears: Their Biology and Management 7: 75–82.

Nakamura, S., Yokoyama, M., Katayama, A., Morimitsu, Y. & Saita, E. (2011). Growth pattern and its characteristics of outer morphology of Asiatic black bear. Wildlife Monograph of Hyogo 3: 107–116.

Liu, D., Wei, R., Zhang, G., et al. (2008). Male panda (Ailuropoda melanoleuca) urine contains kinship information. Chinese Science Bulletin 53: 2793–2800.

Ngoprasert, D., Reed, D. H., Steinmetz, R. & Gale, G. A. (2012). Density estimation of Asian bears using photographic capture– recapture sampling based on chest marks. Ursus 23(2): 117–133.

Martin-Wintle, M. S., Shepherdson, D., Zhang, G., et al. (2015). Free mate choice enhances conservation breeding in the endangered giant panda. Nature Communications 6: 10125.

Nie, Y.-G., Zhang, Z.-J., Swaisgood, R. R. & Wei F.-W. (2012a). Effects of season and social interaction on fecal testosterone metabolites in wild male giant pandas: implications for energetics and mating strategies. European Journal of Wildlife Research 58: 235–241.

Phillips, W. W. A. (1984). Manual of the mammals of Sri Lanka (pp. 290–296). Wildlife and Nature Protection Society of Sri Lanka.

Nie, Y., Swaisgood, R. R., Zhang, Z., Liu, X. & Wie, F. (2012b). Reproductive competition and fecal testosterone in wild male giant pandas (Ailuropoda melanoleuca). Behavioral Ecology and Sociobiology 66: 721–730.

Puschmann, V. W., Schuppel, K. F. & Kronberger, H. (1977). Detection of blastocyst in uterine lumen of Indian bear (Melursus u. ursinus). In: Ippen, R. & Schrader, H. D. (Eds.) Sickness in zoos (pp. 389–391). Berlin: Akademie Verlag.

Martin-Wintle, M. S., Shepherdson, D., Zhang, G., et al. (2017). Do opposites attract? Effects of personality matching in breeding pairs of captive giant pandas on reproductive success. Biological Conservation 207: 27–37. Mauritzen, M., Derocher, A. E. & Wiig, Ø. (2001). Space-use strategies of female polar bears in a dynamic sea ice habitat. Canadian Journal of Zoology 79: 1704–1713. McLellan, B. N. (2005). Sexually selected infanticide in grizzly bears: the effects of hunting on cub survival. Ursus 16(2): 141–156. McLellan, B. N. (2015). Some mechanisms underlying variation in vital rates of grizzly bears on a multiple use landscape. Journal of Wildlife Management 79: 749–765. McLoughlin, P. D., Ferguson, S. H. & Messier, F. (2000). Intraspecific variation

Nie, Y., Speakman, J. R., Wu, Q., et al. (2015). Exceptionally low daily energy expenditure in the bamboo-eating giant panda. Science 349: 171–174. Norris, T. (1969). Ceylon sloth bear. International Wildlife 12: 300–303. Oi, T., Ohnishi, N., Yamada, F. & Kitahara, E. (2008). Characteristics of sex-age composition of Asian black bears killed as nuisances in Kyoto. Mammalian Science 48: 17–24. Okano, T., Nakamura, S., Nakashita, R., et al. (2006). Incidence of ovulation without coital stimuli in captive Japanese black

Penteriani, V., Delgado, M. D. M., Krofel, M., et al. (2018). Evolutionary and ecological traps for brown bears Ursus arctos in human-modified landscapes. Mammal Review 48: 180–193. Peyton, B. (1980). Ecology, distribution, and food habits of spectacled bears, Tremarctos ornatus, in Peru. Journal of Mammalogy 61: 639–652.

Prater S. H. (1965). The book of Indian animals. Bombay: Bombay Natural History Society.

Ramsay, M. A. & Stirling, I. (1986). On the mating system of polar bears. Canadian Journal of Zoology 64: 2142–2151. Ramsay, M. A. & Stirling, I. (1988). Reproductive biology and ecology of female polar bears (Ursus maritimus). Journal of Zoology 214: 601–634. Ratnayeke, S., van Manen, F. T. & Padmalal, U. (2007). Home ranges and habitat use of sloth bears Melursus ursinus inornatus in Wasgomuwa National Park, Sri Lanka. Wildlife Biology 13: 272–284.

33

Systematics, Ecology, and Behavior

Richardson, E. S. (2014). The mating system and life history of the polar bear. PhD dissertation, University of Alberta, Edmonton. Rosell, F., Jojola, S. M., Ingdal, K., et al. (2011). Brown bears possess anal sacs and secretions may code for sex. Journal of Zoology 283: 143–152. Rosenthal, M. & Haggerty, C. (1989). Proceedings of the first international symposium on the spectacled bear. Chicago: Lincoln Park Zoo. Rosenthal, R. A. (1999). International studbook for the spectacled bear (Tremarctos ornatus). Chicago: The Lincoln Park Zoological Society. Rosing-Asvid, A., Born, E. W. & Kingsley, M. C. S. (2002). Age at sexual maturity of males and timing of the mating season of polar bears (Ursus maritimus) in Greenland. Polar Biology 25: 878–883. Schaller, G. B., Hu, J., Pan, W. & Zhu, J. (1985). The giant pandas of wolong. Chicago: University of Chicago Press. Schwartz, C. C., Keating, K. A., Reynolds, H. V., et al. (2003). Reproductive maturation and senescence in the female brown bear. Ursus 14(2): 109–119. Schwartz, C. C., Haroldson, M. A., White, G. C., et al. (2006). Temporal, spatial, and environmental influences on the demographics of grizzly bears in the greater Yellowstone Ecosystem. Wildlife Monographs 161: 1–68. Schwarzenberger, F., Fredriksson, G., Schaller, K. & Kolter, L. (2004). Fecal steroid analysis for monitoring reproduction in the sun bear (Helarctos malayanus). Theriogenology 62: 1677–1692. Sergiel, A., Naves, J., Kujawski, P., et al. (2017). Histological, chemical and behavioural evidence of pedal communication in brown bears. Scientific Reports 7: 1052. Shimozuru, M., Shirane, Y., Tsuruga H., et al. (2019). Incidence of multiple paternity and inbreeding in high-density brown bear populations on the Shiretoko Peninsula, Hokkaido, Japan. Journal of Heredity 110(3): 321–331. Smith, T. G. & Aars, J. (2015). Polar bears (Ursus maritimus) mating during late June on the pack ice of northern Svalbard, Norway. Polar Research 34: 25786. Spady, T. J., Lindburg, D. G. & Durrant, B. S. (2007). Evolution of reproductive

34

seasonality in bears. Mammal Review 37: 21–53. Stenhouse, G., Boulanger, J., Lee, J., et al. (2005). Grizzly bear associations along the eastern slopes of Alberta. Ursus 16(1): 31–40. Steyaert, S. M. J. G., Endrestøl, A., Hackländer, K., Swenson J. E. & Zedrosser, A. (2012). The mating system of the brown bear Ursus arctos. Mammal Review 42: 12–34. Steyaert, S. M. J. G., Swenson, J. E. & Zedrosser, A. (2014). Litter loss triggers estrus in a nonsocial seasonal breeder. Ecology and Evolution 4: 300–310. Steyaert, S. M. J. G., Leclerc M., Pelletier, F., et al. (2016). Human shields mediate sexual conflict in a top predator. Proceedings of the Royal Society B: Biological Sciences 283: 20160906. Stirling, I., Spencer, C. & Andriashek, D. (2016). Behavior and activity budgets of wild breeding polar bears (Ursus maritimus). Marine Mammal Science 32: 13–37. Stone, I. R. & Derocher, A. E. (2007). An incident of polar bear infanticide and cannibalism on Phippsoya, Svalbard. Polar Record 43: 171–173. Swaisgood, R. R., Lindburg, D. G. & Zhang, H. (2002). Discrimination of oestrous status in giant pandas (Ailuropoda melanoleuca) via chemical cues in urine. Journal of Zoology 257: 381–386. Swenson, J. E., Sandegren, F., Söderberg, A., et al. (1997). Infanticide caused by hunting of male bears. Nature 386: 450–451. Tattoni, C., Bragalanti, N., Groff, C. & Rovero, F. (2015). Patterns in the use of rub trees by the Eurasian brown bear. Hystrix, the Italian Journal of Mammalogy 26: 118–124. Tibbetts, E. A. & Dale, J. (2007). Individual recognition: it is good to be different. Trends in Ecology and Evolution 22: 529–537.

yesoensis). Journal of Reproduction and Development 39: 325–331. Tsubota, T., Howell-Skalla, L., Boone, W. R., Garshelis, D. L. & Bahr, J. M. (1998). Serum progesterone, oestradiol, luteinizing hormone and prolactin profiles in the female black bear (Ursus americanus). Animal Reproduction Science 53: 107–118. Vaisefeld, M. A. & Chestin, I. E. (1993). Bears: brown bear, polar bear, Asian black bear. Distribution, ecology, use and protection. Moscow: Nauka. Wei, W., Swaisgood, R. R., Owen, M. A., et al. (2019). The role of den quality in giant panda conservation. Biological Conservation 231: 1892196. Wiig, Ø., Gjertz, I., Hansson, R. & Thomassen, J. (1992). Breeding behaviour of polar bears in Hornsund, Svalbard. Polar Record 28: 157–159. Wong, S., Servheen, C., Ambu, L. & Norhayati, A. (2005). Impacts of fruit production cycles on Malayan sun bears and bearded pigs in lowland tropical forest of Sabah, Malaysian Borneo. Journal of Tropical Ecology 21: 627–639. Yamamoto, K., Tsubota, T. & Kita, I. (1998). Observation of sexual behavior of captive Japanese black bears, Ursus tibetanus japonicus. Journal of Reproduction and Development 44: j13–j18. Yamamoto, T., Tamatani, H., Tanaka, J., et al. (2012). Multiple paternity in Asian black bear Ursus thibetanus (Ursidae, Carnivora) determined by microsatellite analysis. Mammalia 77: 215–217. Yamane, M., Yamamoto, Y., Tsujimoto, T. & Osawa, T. (2009). Relationship between uterine morphology and peripheral concentrations of sex steroid hormone in wild Japanese black bears (Ursus thibetanus japonicus). Animal Reproduction Science 113: 251–262. Yamazaki, K. (2017). Challenges in management and conservation of Asian black bears in Japan. Tokyo: University of Tokyo Press.

Trivers, R. L. (1972). Parental investment and sexual selection. In: Campbell B. G. (Ed.), Sexual selection and the descent of man 1871–1971 (pp. 136–179). Chicago: Aldine.

Yoganand, K. (2005). Behavioural ecology of sloth bear (Melursus ursinus) in Panna National Park, Central India. PhD Thesis. Saurashtra University, India.

Tsubota, T. & Kanagawa, H. (1993). Morphological characteristics of the ovary, uterus and embryo during the delayed implantation period in the Hokkaido brown bear (Ursus arctos

Yoganand, K., Rice, C. G. & Johnsingh, A. J. T. (2013). Sloth bear: Melursus ursinus. In: Johnsingh, A. J. T. & Manjrekar, N. (Eds.), Mammals of South Asia (pp. 438–456). Telangana, India: Orient Blackswan pvt. Ltd.

Mating Strategies

Zedrosser, A., Bellemain, E., Taberlet, P. & Swenson, J. E. (2007). Genetic estimates of annual reproductive success in male brown bears: the effects of body size, age, internal relatedness and population density. Journal of Animal Ecology 76: 368–375. Zeyl, E., Aars, J., Ehrich, D., Bachmann L. & Wiig, Ø. (2009). The mating system of polar bears: a genetic approach. Canadian Journal of Zoology 87: 1195–1209. Zhang, H., Li, D., Wang, C. & Hull, V. (2009). Delayed implantation in giant

pandas: the first comprehensive empirical evidence. Reproduction 138: 979–986. Zhang, Z., Swaisgood, R. R., Zhang, S., et al. (2011). Old-growth forest is what giant pandas really need. Biology Letters 7(3): 403–406. Zhang, Z., Sheppard, J. K., Swaisgood, R. R., et al. (2014). Ecological scale and seasonal heterogeneity in the spatial behaviors of giant pandas. Integrative Zoology 9: 46–60.

Zhu, L., Wu, Q., Dai, J., Zhang, S. & Wei, F. (2011). Evidence of cellulose metabolism by the giant panda gut microbiome. Proceedings of the National Academy of Sciences 108: 17714–17719. Zhu, X., Lindburg, D. G., Pan, W., Forney, K. A. & Wang, D. (2001). The reproductive strategy of giant pandas: infant growth and development and mother-infant relationships. Journal of Zoology 253: 141–155.

35

Chapter

3

Interspecific Interactions between Brown Bears, Ungulates, and Other Large Carnivores Andre´s Ordiz, Miha Krofel, Cyril Milleret, Ivan Seryodkin, Aimee Tallian, Ole-Gunnar Støen, Therese Ramberg Sivertsen, Jonas Kindberg, Petter Wabakken, Ha˚kan Sand, and Jon E. Swenson

Introduction Large carnivores, such as brown bears (Ursus arctos), wolves (Canis lupus), and tigers (Panthera tigris), can play a key ecological role from their apex position in trophic systems (Ordiz et al. 2013). Within the overall context of bottom-up and top-down regulation of ecosystems (Sinclair & Krebs 2002), predation by large carnivores often induces demographic and behavioral changes in prey species. These vertical interactions between different trophic levels are important regulatory mechanisms in nature. On the other hand, competitive interactions between species, or horizontal interactions within the same trophic level, are also common. Interspecific interactions between large carnivores are widespread in many ecosystems and can play an important role in community structure and stability. For instance, competition between carnivores can reduce the population size of their respective competitor, which in turn affects lower trophic levels (Palomares & Caro 1999; Caro & Stoner 2003). Competition and other interactions among carnivores can lead to spatial avoidance, shifts in habitat use, altered predation patterns, and other behavioral changes. For example, kleptoparasitism or the theft of kills can force affected species to additional hunting (Krofel et al. 2012; Elbroch et al. 2015). Indirect competition for shared resources, such as prey or habitat, can also affect population dynamics, predation patterns, and habitat selection (e.g. Ordiz et al. 2015). Such interactions can either weaken or strengthen the top-down effects of carnivores by altering predator densities or predation patterns. For instance, scavenging, with kleptoparasitism between carnivores as one of its major mechanisms, is a key ecological process involved in energy flow in ecosystems (DeVault et al. 2003). Interspecific interactions can cause the extirpation of one species, but more often result in resource partitioning and, thus, coexistence (Apps et al. 2006). The outcome of interactions may be mediated by the environmental context and the effects of humans, which can have an overwhelming impact on the population dynamics of the interacting species, their distribution, and the effects of competition (Apps et al. 2006; Krofel & Jerina 2016). In fact, this may be the norm for large carnivores inhabiting human-dominated landscapes, given the long history of human persecution and the general avoidance of people by large carnivores (see Chapter 18). Overall,

36

predator–prey relationships and other interspecific interactions are important topics for ecology and management, because they can influence the distribution and abundance of species (Schoener 1983; Wisz et al. 2013). Because interactions between apex predators can have cascading effects on lower trophic levels, research and management would benefit from a multispecies or ecosystem-focused approach, rather than focusing efforts on single species (Ordiz et al. 2015; Krofel & Jerina 2016). In addition, large carnivores generate considerable public interest and are a central issue in human–wildlife conflicts. Predation is the mechanism driving apex predators’ function in nature, but it is also a source of conflict with different stakeholders, e.g. hunters and livestock owners, when predation affects wild game, domestic or semidomestic species (depredation). This situation is challenging when trying to secure long-term carnivore conservation and coexistence with people in the human-dominated landscapes that currently characterize most of our planet. Brown bears are very efficient predators of newborn ungulates. Predation by brown bears often results in additive neonate ungulate mortality (e.g. Griffin et al. 2011, see below), which is of particular interest in ecosystems where bears are sympatric with other large carnivores, such as canids and felids, that also prey on ungulates. This is especially relevant given the great extent to which brown bears co-occur with other carnivores in the Northern Hemisphere (Figure 3.1). In this chapter, we summarize the effects of brown bear predation on both wild and domestic ungulate populations and present the current state of knowledge on interactions between brown bears and other large carnivores, focusing on felids and canids (in particular, wolves). Besides the inherent interest of this topic from an ecological perspective, it also has direct implications for the conservation and management of predators and their ungulate prey; e.g. the latter are both commonly exposed to human management and predation by large carnivores (Jonzén et al. 2013).

Bears and Wild Ungulate Populations All bear species have the potential to kill wild ungulates, but because bears are omnivores and opportunistic predators, they have relatively loose predator–prey relationships with ungulates. This means that bear population densities are probably

Interspecific Interactions

Ursus arctos Ursus arctos with Lynx lynx (in Eurasia) Ursus arctos with Lynx canadensis (in North America)

Ursus arctos Ursus arctos with Panthera tigris (Asian Far East) Ursus arctos with Panthera pardus (central Asia) Ursus arctos with Puma concolor Ursus arctos with Panthera pardus and Panthera tigris (Asian Far East) Ursus arctos with Panthera uncia (central Asia)

Ursus arctos Ursus arctos with Canis lupus

Figure 3.1 Worldwide distribution of brown bears and its overlap with the distribution of several felids (top and central panel) and wolves (bottom panel) in the Northern Hemisphere. (Source: IUCN 2019. The IUCN Red List of Threatened Species, Version 6.2, www.iucnredlist.org. Downloaded March 16, 2019.) (A black and white version of this figure will appear in some formats. For the color version, please refer to the plate section.)

37

Systematics, Ecology, and Behavior

Figure 3.2 Moose calf (top panel) and reindeer calf (bottom panel) killed by brown bears in Sweden. (Photos by A. Ordiz.)

more affected by the abundance of other food sources than ungulates, i.e. bear densities may fluctuate independently of ungulate densities. However, the effect of bear predation on an ungulate population may primarily vary with changes in the relative density of the bear and ungulate species, i.e. the effect of bears is likely stronger as their relative density increases (Zager & Beecham 2006). Brown bears are efficient predators on neonatal moose (Alces alces), elk (Cervus canadensis), caribou (reindeer in Europe; Rangifer tarandus), and muskoxen (Ovibos moschatus) in North America (Zager & Beecham 2006; Arthur & Vecchio 2017) and neonatal moose and reindeer in Europe (Swenson et al. 2007; Karlsson et al. 2012; Figure 3.2). Although less common, bears can also prey on adult moose, elk, caribou, and muskoxen in North America (Zager & Beecham 2006; Arthur & Vecchio 2017) and on adult moose, red deer (Cervus elaphus), and wild boar (Sus scrofa) in Europe (Dahle et al.

38

2013; Krofel et al., unpublished data). Bears apparently do not regularly prey on American bison (Bison bison) (Wyman 2002) or sika deer (Cervus nippon) (Sato et al. 2005). It is worth mentioning that other bear species are also predatory. American black bears (Ursus americanus) are efficient predators of neonatal moose, elk, caribou, white-tailed deer (Odocoileus virginianus), and mule deer (Odocoileus hemionus), but very rarely prey on adult ungulates (Zager & Beecham 2006; Monteith et al. 2014). In Taiwan, the relatively small-sized adult Formosan muntjac (Muntiacus reevesi) and Formosan serow (Naemorhedus swinhoei) are part of the diet of Asiatic black bears (Hwang et al. 2002), which also prey on Japanese serow (Capricornis crispus) (Huygens et al. 2003). Although Andean bears have been reported to attack woolly tapir (Tapirus pinchaque) and prey on white-tailed deer, guanaco (Lama guanicoe), and vicuña (Vicugna vicugna), there are few supporting quantitative data (Stirling & Derocher 1990). Polar bears (Ursus maritimus) scavenge carcasses when they find them, and occasionally kill reindeer (Derocher et al. 2000). The effects of bear predation on wild ungulate populations vary enormously, but predation generally appears to be additive at low ungulate densities and more compensatory as ungulate population densities approach carrying capacity (Zager & Beecham 2006). Compensatory mortality means that instead of causing additional mortality to the population, as would be the case with additive mortality, the occurrence of predation does not cause higher total mortality rates. Additionally, the overall impact of bear predation on ungulate population growth is dependent on the targeted segment of the prey population, e.g. variation in age- and sex-specific survival differentially affect population growth (Gervasi et al. 2012). In seven of eight reviewed studies of moose populations in Alaska, brown and American black bear predation dominated calf mortality and this mortality was primarily additive (Boertje et al. 2010). These findings are consistent with studies in other brown bear–moose systems, such as in Scandinavia (Swenson et al. 2007), and other bear–ungulate systems in North America (Griffin et al. 2011). Bear predation on neonates, especially in combination with wolf predation, has probably caused negative population growth in several populations of moose and elk (Garrott et al. 2009; Boertje et al. 2010). Brown bear predation on adult ungulates is seemingly compensatory to other sources of mortality; however, exceptions have been observed in several systems (e.g. Zager & Beecham 2006). For example, brown bear predation on adult and calf muskoxen apparently increased in an area of Alaska, perhaps because caribou and moose populations had declined, causing a decline in the muskox population (Arthur & Vecchio 2017). Nevertheless, Keech et al. (2000) in Alaska and Swenson et al. (2007) in Scandinavia showed the existence of reproductive compensation in moose females following loss of calves. Female moose can have higher reproductive rates the year after predation occurred, i.e. managers should expect partial compensation of large carnivore predation through higher reproduction the

Interspecific Interactions

year after a moose female lost her calves, especially if this happened early in the calves’ lives (Swenson et al. 2007).

Bears and Domestic Ungulate Populations: A Case Study from Scandinavia Brown bears prey on domestic and semi-domestic livestock in many parts of their range, including sheep, cows, goats, and reindeer (e.g. Knight & Judd 1979; Kaczensky 1999). Here, we highlight the case of brown bear predation of semi-domestic reindeer in northern Scandinavia as a case study of human–bear conflict. Brown bears can be efficient depredators of semidomestic reindeer, especially during calving in spring, when neonates in their first weeks of life are easy prey for bears (Karlsson et al. 2012; Sivertsen 2017). Large losses of reindeer calves to bear predation frequently lead to conflict with reindeer owners. Pastoral reindeer husbandry, with more than 1200 years of tradition in Fennoscandia (Ruong 1982), is an essential part of the culture and livelihoods of the Sámi people (Zabel & Holm-Müller 2008). Traditionally, the reindeer owners have protected their reindeer by killing bears on their calving grounds. However, the core areas of traditional Sámi reindeer herding were also the main areas where bears avoided extirpation in Sweden during the government-driven large carnivore extermination in the nineteenth century. In 1930, only ~130 individual bears survived in mountainous areas in northern and central Sweden and one area in Norway (Swenson et al. 1995). After the abolishment of bounties in 1893 and legal protection in 1986, the bear population increased and reached ~3000 bears in Sweden in 2008 (Kindberg et al. 2011; Swenson et al. 2017). This bear population is once again decreasing, due to increased harvest quotas set by management agencies, mostly driven by conflicts with reindeer husbandry and moose hunters, in order to stop the population growth. Although reindeer herders claim that losses are substantial, the absence of reliable data on brown bear predation has made it difficult to reach an agreement on compensation schemes and preventive interventions safeguarding both Sámi reindeer herding and a viable brown bear population. In order to address the need for reliable data, the Swedish government initiated a study on brown bear kill rate on reindeer in forested calving grounds in 2010–2012 (Karlsson et al. 2012). The results showed that brown bears killed calves almost exclusively during, and shortly after, peak calving, i.e. in the three last weeks of May. The estimated predation rate indicated that brown bears may have caused a considerable proportion (39% and 67%) of the total calf losses observed within the two reindeer herding districts. Average annual calf mortality in the herding districts was approximately 43% and 41%, and the estimated annual bear-caused calf mortality was around 29% and 16% of the total number of calves born annually within the two studied reindeer herding districts (Sivertsen 2017). A follow-up study on interventions in 2013–2016 showed that calving in corrals prevents most losses to bears, but was too expensive, increased the risk of diseases, and was not compatible with traditional

Sámi reindeer herding traditions (Frank et al. 2017). Reduction of the bear density in calving grounds, either as management killing or area-differentiated license hunting, was considered the most efficient way to solve the problem (Frank et al. 2017). Nevertheless, a zoning system, prioritizing carnivore conservation and reindeer herding in different areas, might also help reduce conflict (Ordiz et al. 2017). The recent increase in hunting quotas has given hunters the opportunity to reduce the bear population, but increased bear harvest does not seem to have reduced the conflict in the reindeer husbandry areas. This is because calving areas are remote, far from roads, and less available for brown bear hunters. Therefore, most of the removal of bears in calving areas has been carried out as authorized spring management culls (i.e. not as part of the regular hunting season), often targeting females with cubs and using helicopters. This has raised questions about the ethics of management removals and caused conflict among stakeholders, including the general public, people interested in bear conservation, and hunter organizations.

Brown Bear Interactions with Other Large Carnivores Brown bears interact with many other large carnivores, including other bear species (e.g. Schwartz et al. 2010), felids (e.g. Krofel et al. 2012), and canids (e.g. Ordiz et al. 2015). There is extensive overlap in the distribution of brown bears, several felids, and wolves, the largest canid species, throughout the Northern Hemisphere (Figure 3.1). In particular, this section focuses on brown bear–felid and brown bear–wolf interactions, while acknowledging that interactions with several bear species and other canids and felids also occur where their distributions overlap. Much of the existing literature on interactions between large carnivores reports on direct and indirect interactions at the individual level. Yet, we still lack knowledge about the effects of interspecific interactions beyond individuals, i.e. at the population level. Scarcity of such information has been specifically highlighted for the interactions between wolves and bears (Ballard et al. 2003) and, more broadly, for large carnivores in general (Périquet et al. 2015). In fact, interspecific competition between free-ranging apex predators is challenging to quantify, because such species are elusive and logistically complex and expensive to study. Interactions between bears and felids or canids include all the forms of interspecific interactions described earlier in the chapter, i.e. they use the kills of each other via kleptoparasitism, they can displace each other and, in agonistic situations, they can kill each other (see below). As the largest terrestrial scavengers with excellent olfactory abilities, brown bears are indeed one of the most important dominant scavengers and kleptoparasites in the Holarctic region (Krofel et al. 2012), and kleptoparasitism is likely the most common form of interaction between bears and other carnivores.

39

Systematics, Ecology, and Behavior

Brown Bears and Felids Throughout their range in the Northern Hemisphere, brown bears overlap with multiple species of wild cats. However, the available literature on interspecific interactions is limited to tiger, mountain lion (Puma concolor), Eurasian lynx (Lynx lynx), common leopard (Panthera pardus), and snow leopard (Panthera uncia) (Figures 3.1 and 3.3). All of them prey primarily on wild ungulates, and kleptoparasitism is the most frequently reported interaction between brown bears and felids (Krofel et al. 2012). Smaller felids predominantly hunt small prey (i.e. smaller than their body size), which are quickly consumed, and this probably explains their low level of interactions with bears. On the other hand, larger felids mainly hunt ungulates, which may take them several days to consume. Extended handling time increases the potential for carcasses to be usurped by the scavenging bears (Krofel et al. 2012; Allen et al. 2015). It has also been suggested that scavenging by bears may have been an important factor in shaping felid patterns of prey selection toward smaller prey and results in consequent morphological and behavioral adaptations of predators during the evolution of some felids, like the genus Lynx (Krofel et al. 2012). In the Rocky Mountains of North America, brown bears and American black bears frequently scavenge on mountain lion kills, providing them with an important source of energy (Murphy et al. 1998; Elbroch et al. 2015). In that system, bears displaced mountain lions from 7% of their kills in winter and 12% of kills between spring and autumn (Murphy et al. 1998). Similarly, brown bears usurp ungulate carcasses killed by Eurasian lynx in Asia and Europe (Zhiryakov & Baidavletov 2003; Mattisson et al. 2011; Krofel et al. 2012), sometimes by following lynx tracks in the snow (Kostoglod 1981; Krofel & Kos 2007). Kleptoparasitic interactions between brown bears and Eurasian lynx were studied in detail in the Dinaric Mountains of Slovenia and Croatia, where the predicted proportion of all lynx kills found by bears throughout the year was 32% and lynx were estimated to lose 15% of the biomass of their ungulate kills. In response, lynx increased ungulate kill rate by 23%, but due to long searching times needed to capture new prey, they were only able to compensate for 59% of the lost biomass (Krofel et al. 2012). The frequency of bear scavenging on lynx kills was strongly dependent on bear density and activity patterns. Theft of kills was also highest during lynx pregnancy and lactation periods, suggesting that kleptoparasitism by bears could affect lynx reproductive success (Krofel et al. 2012). In Slovenia, these interactions were strongly influenced by bear management practices. Kleptoparasitism intensity differed almost threefold between different bear management zones (e.g. interaction increased in zones with higher bear protection level and more intensive artificial feeding) and the presence of an artificial bear feeding site locally increased the odds of lynx losing kills fivefold (Krofel & Jerina 2016). A bear was reported consuming a lynx in Kamchatka, but also a lynx was observed feeding on a bear

40

Figure 3.3 Scat of brown bear with tail and hair of tiger (top) and tiger carcass partially consumed by a brown bear (bottom). (Photos by I. Seryodkin.)

carcass in the same area (Mosolov & Valentsev 2003). However, such observations could also represent scavenging rather than predation events. Although brown bears are typically dominant in interactions with lynx and mountain lions, the situation is less

Interspecific Interactions

Figure 3.4 Location of wolf pairs (black dots) and brown bear density (grayscale shadows show the densest bear areas) in south-central Scandinavia, northern Europe, during the ongoing recovery phase of the wolf population, in 1985 (left), 2000 (central), and 2010 (right panel). Modified from Ordiz et al. (2015).

predictable in interactions with tigers. Among 45 documented cases of aggressive encounters between tigers and brown bears, the fight ended with the death of the bear in 51% of the cases, of the tiger in 27%, while both animals survived in 22% (Seryodkin et al. 2018). According to tiger diet analyses, brown bears can account for up to 5% of a tigers’ diet, although Asiatic black bears are more commonly consumed (Tkachenko 2012). Consumption does not necessarily mean predation, as tigers also scavenge on bears that died from other causes. Although tigers kill most bears during their active period, tigers also predate on hibernating bears in their dens. Most of the brown bears killed by tigers are young animals, or females, and are generally killed by male tigers (Seryodkin et al. 2018). Predation on tigers by brown bears is probably more intensive in years with lower food availability, when bears were occasionally reported to chase tigers for long distances (Kostoglod 1981). Predation by brown bears also contributes to tiger cub mortality (Gorokhov 1997). Despite the risk, bears will also consume the remains of tiger-killed prey. In the Russian Far East, signs of kleptoparasitism by brown or Asiatic black bears were recorded at 17% of tiger kills. In cases where brown bears were present at tiger kills, bears either used the carcass after the tigers already left (42%), usurped the kill from the tiger (25%), or both species shared the kill by alternately feeding on it (17%). For the remaining 17% of the cases, the order of use was unknown. Most of the tigers in these cases were females (Seryodkin et al. 2018). On rare occasions, tigers scavenge on bear-killed prey. For example, Rakov (1965) reported that a bear killed an adult tigress while it was eating a wild boar killed by that bear. Finally, interactions between brown bears and common and snow leopards remain mostly unstudied. Pikunov and Korkishko (1992) reported a case from the Far East where a brown bear followed the tracks of a leopard and usurped its kill. Heptner and Sludskii (1972) mentioned the single case of a two-year-old brown bear killed and partly eaten by a snow leopard in the Tien Shan range.

Brown Bears and Canids Brown bear populations overlap with several canid species throughout the Northern Hemisphere. For instance, coyotes (Canis latrans) overlap the entire brown bear distribution in North America, and bears overlap with golden jackals (Canis aureus) in different parts of Eurasia. However, the red fox (Vulpes vulpes) and the wolf have the greatest overlap with brown bears; both are sympatric with bears across most of their range in North America and Eurasia. Here, we focus on interactions between bears and wolves (Figure 3.1), which have received greater attention than interactions between bears and smaller canids. Similar to bear–felid interactions, bears often kleptoparasitize wolf kills, both in North America (Ballard et al. 2003; Smith et al. 2003) and Europe (Milleret 2011; Ordiz et al. 2015). The outcome of interactions between bears and wolves at carcasses varies, but in general, bears often dominate and limit the access of wolves to kills (Boertje et al. 1988; MacNulty et al. 2001). For instance, Ballard et al. (2003) compiled 108 direct bear–wolf interactions in North America, outside Yellowstone National Park, and 57% of them occurred at kill sites, with bears dominating in most of those encounters (Ballard et al. 2003). In Europe, brown bears were recorded at 20% of wolf-killed prey in Slovenia, and 50–60% of wolf kills in Scandinavian studies, but wolves were never recorded at bear-killed prey (authors’ unpublished data). Nevertheless, simultaneous scavenging by both species has also been reported (Smith et al. 2003; Lewis & Lafferty 2014). Beyond interactions at specific locations, such as kill sites or dens of brown bears and wolves (Ballard et al. 2003), the potential effect of interspecific interactions between these large carnivores has been studied in Scandinavia at larger scale in recent years. The recent and ongoing recovery of the wolf population in areas where brown bears have been both continually present and absent in Sweden and Norway, in combination with long-term monitoring efforts of both species,

41

Systematics, Ecology, and Behavior

offered a great opportunity to understand competition and coexistence between these two apex predators. Bears and wolves heavily rely on a common prey resource, i.e. seasonally available moose calves (Swenson et al. 2007; Sand et al. 2008), which sets the stage for competition. For these reasons, much of what we know about the nature of interactions between wolves and bears comes from Scandinavia, where researchers are studying bear–wolf interactions with two complementary approaches, in terms of habitat selection and kill rates. At the landscape level and during the ongoing demographic recovery and expansion of the wolf population in Scandinavia, the establishment of wolf pairs was negatively affected by the presence of bears and humans (Ordiz et al. 2015; Figure 3.4). At the finer scale of habitat selection, inside home ranges of radio-collared bears and wolves and within the area where both species overlap in central Scandinavia, bears and wolves segregated more than expected by chance (Milleret et al. 2018). Wolves tended to select for moose occurrence, young forests, and rugged terrain more than bears, which likely reflected the different requirements of an obligate carnivore and an omnivore. However, both species generally avoided human-related habitats during daytime (Milleret et al. 2018). Regarding predation rates, it was previously suggested that wolf kill rates would be higher where they coexisted with brown bears, because they would be forced to hunt more often to compensate for the loss of food caused by bear kleptoparasitism (e.g. Ballard et al. 2003). However, recent research suggests that wolves sympatric with bear populations may actually kill less often. A study comparing wolf kill rates with respect to bear presence–absence in Scandinavia and Yellowstone National Park found that bear presence was negatively correlated with wolf kill rate in both systems (Tallian et al. 2017). The mechanism behind this pattern is under investigation, but it could involve either interference competition, i.e. a species directly alters the resource-attaining behavior of another species, or exploitation competition, i.e. species interact indirectly competing for common resources, such as territory or food, or both. Bear presence may increase wolf handling time (e.g. wolves stay longer at a kill to defend or gain access to the carcass), search time (e.g. wolf and bear predation both reduce the availability of shared prey, making them more difficult to find), or both. Regardless of the

References Allen, M. L., Elbroch, L. M., Wilmers, C. C. & Wittmer, H. U. (2015). The comparative effects of large carnivores on the acquisition of carrion by scavengers. The American Naturalist 185: 822–833. Apps, C. D., McLellan, B. N. & Woods, J. G. (2006). Landscape partitioning and spatial inferences of competition between black and grizzly bears. Ecography 29: 561–572.

42

mechanism explaining these results, they suggest that the total impact of wolves and brown bears on prey may be less than the sum of their individual impacts. If so, interactions between wolves and bears may partially dampen the effect of the other carnivore on ungulate mortality (Tallian et al. 2017). Habitat- and predation-related studies (Ordiz et al. 2015; Tallian et al. 2017; Sanz-Pérez et al. 2018) suggest that brown bears have a negative impact on wolf habitat selection and foraging ability, such that wolf populations that coexist with brown bears may suffer fitness consequences. However, wolf presence may have the opposite effect on bear populations. Wolf kills are an important source of protein for scavenger communities, including brown and American black bears (Wilmers et al. 2003). Wolves kill ungulates consistently throughout the course of a year, providing a spatiotemporal stable source of meat for bears and other scavengers (Wikenros et al. 2013). Consistent access to wolf kills may even provide a buffer to climate change-induced reductions in winter-killed ungulates. Shorter winters may lower ungulate mortality, with fewer carcasses available for scavengers, and this is an important source of post-denning protein for bears in some systems (Wilmers & Getz 2005). Ultimately, determining the energetic costs of interactions between bears and other large carnivores, and ascertaining their consequences for predator population dynamics, will help us understand the costs of sympatry among apex predator populations. Understanding the costs of sympatry will allow us to clarify the effect of carnivore coexistence on prey populations and lower trophic levels. Gaining detailed information on coexisting apex predators will also help tune the adaptive management of economically important ungulate game species, such as moose, elk or red deer, and other deer species. In this sense, the current recovery of some large carnivores in Europe (Chapron et al. 2014) and North America (Bruskotter & Shelby 2010), along with the expanded use of GPS-based telemetry, should provide opportunities to study interspecific competition among sympatric apex predators at both the individual and population level across multiple ecosystems. Brown bears co-occur with multiple large carnivores (Figure 3.1) and ungulate species throughout their current distribution in the Northern Hemisphere, which highlights the importance of understanding the outcomes of interspecific interactions among carnivores and with their prey populations.

Arthur, S.M. & Vecchio, P.A. Del. (2017). Effects of grizzly bear predation on muskoxen in northeastern Alaska. Ursus 28: 81–91. Ballard, W., Carbyn, L. N. & Smith, D. (2003). Wolf interactions with nonprey. In: Mech, L. D. & Boitani, L. (Eds.), Wolves: Behavior, ecology, and conservation (pp. 259–271). Chicago, IL: University of Chicago Press.

Boertje, R., Keech, M. & Paragi, T. (2010). Science and values influencing predator control for Alaska moose management. Journal of Wildlife Management 74: 917–928. Boertje, R. D., Gasaway, W. C., Grangaard, D. V. & Kelleyhouse, D. G. (1988). Predation on moose and caribou by radio-collared grizzly bears in east central Alaska. Canadian Journal of Zoology 66: 2492–2499.

Interspecific Interactions

Bruskotter, J. T. & Shelby, L. B. (2010). Human dimensions of large carnivore conservation and management. Human Dimensions of Wildlife 15: 311–314. Caro, T. M. & Stoner, C .J. (2003). The potential for interspecific competition among African carnivores. Biological Conservation 110: 67–75. Chapron, G., Kaczensky, P., von Arx, M., et al. (2014). Recovery of large carnivores in Europe’s modern human-dominated landscapes. Science 346: 1517–1519. Dahle, B., Wallin, K., Cederlund, G., et al. (2013). Predation on adult moose Alces alces by European brown bears. Wildlife Biology 19: 165–169. Derocher, A. E., Wiig, Ø. & Bangjord, G. (2000). Predation of Svalbard reindeer by polar bears. Polar Biology 23: 675–678. DeVault, T. L., Rhodes, O. E. & Shivik, J. A. (2003). Scavenging by vertebrates: behavioral, ecological, and evolutionary perspectives on an important energy transfer pathway in terrestrial ecosystems. Oikos 102: 225–234. Elbroch, L. M., Lendrum, P. E., Allen, M. L. & Wittmer, H. U. (2015). Nowhere to hide: pumas, black bears, and competition refuges. Behavioral Ecology 26: 247–254. Frank, J., Støen, O. G., Segerström, P., et al. (2017). Kalvning i hägn och områdesriktad jakt på björn som åtgärder för att minska björnars predation på ren. Rapport från Viltskadecenter 7. Sveriges Lantbruksuniversitet SLU, Grimsø forskningsstation, Riddarhyttan, Sweden. [In Swedish.] https://pub.epsilon.slu.se/ 15369/7/frank_j_et_al_180425.pdf. Garrott, R. A., White, P. J. & Watson, F. G. R. (2009). The ecology of large mammals in central Yellowstone: Sixteen years of integrated field studies. Terrestrial Ecology Series Volume 3. San Diego, CA: Academic Press. Gervasi, V., Nilsen, E. B., Sand, H., et al. (2012). Predicting the potential demographic impact of predators on their prey: a comparative analysis of two carnivore–ungulate systems in Scandinavia. Journal of Animal Ecology 81: 443–454. Gorokhov, G. F. (1997). The abundance and structure of the Amur tiger population in southern Sikhote-Alin, in Redkie vidy mlekopitayushchikh i ikh okhrana. In: Proceedings 2nd All-Union Conf. “Rare Species of Mammals and Their

Conservation (pp. 119–120). Moscow: Nauka.

endangered apex predator. Biological Conservation 197: 40–46.

Griffin, K. A., Hebblewhite, M., Robinson, H. S., et al. (2011). Neonatal mortality of elk driven by climate, predator phenology and predator community composition. Journal of Animal Ecology 80: 1246–1257.

Krofel, M. & Kos, I. (2007). Evidence of the brown bear (Ursus arctos) tracking the Eurasian lynx (Lynx lynx) on the Snežnik plateau, Slovenia. Natura Sloveniae 9: 45–46.

Heptner, V. G. & Sludskii, A. A. (1972). Mlekopitayushchiye Sovetskogo Soyuza. In Mammals of the Soviet Union, Carnivora. Moscow: Vysshaya Shkola/

Krofel, M., Kos, I. & Jerina, K. (2012). The noble cats and the big bad scavengers: effects of dominant scavengers on solitary predators. Behavioral Ecology and Sociobiology 66: 1297–1304.

Huygens, O. C., Miyashita, T., Dahle, B., Carr, M. & Izumiyama, S. (2003). Diet and feeding habits of Asiatic black bears in the Northern Japanese Alps. Ursus 14: 236–245. Hwang, M.-H., Garshelis, D. L. & Wang, Y. (2002). Diets of Asiatic black bears in Taiwan, with methodological and geographical comparisons. Ursus 13: 111–125. Jonzén, N., Sand, H., Wabakken, P., et al. (2013). Sharing the bounty – adjusting harvest to predator return in the Scandinavian human–wolf–bear–moose system. Ecological Modelling 265: 140–148. Kaczensky, P. (1999). Large carnivore depredation on livestock in Europe. Ursus 11: 59–72. Karlsson, J., Støen, O. G., Segerström, P., et al. (2012). Björnpredation på ren och potentiella effekter av tre förebyggande åtgärder. Rapport från Viltskadecenter 6. Sveriges Lantbruksuniversitet SLU, Grimsø forskningsstation, Riddarhyttan, Sweden. [In Swedish.] Keech, M. A., Bowyer, R. T., Hoef, J. M. Ver, et al. (2000). Life-history consequences of maternal condition in Alaskan moose. The Journal of Wildlife Management 64: 450–462. Kindberg, J., Swenson, J. E., Ericsson, G., et al. (2011). Estimating population size and trends of the Swedish brown bear Ursus arctos population. Wildlife Biology 17: 114–123. Knight, R. R. & Judd, S. (1979). Grizzly bears that kill livestock. International Conference on Bear Research and Management 5: 186–190. Kostoglod V. E. (1981). The experience of prolonged tracking of the brown bear in Sikhote-Alin. Bulletin of Moscow Society of Naturalists. Biological series 86: 3–12. Krofel, M. & Jerina, K. (2016). Mind the cat: conservation management of a protected dominant scavenger indirectly affects an

Lewis, T. M. & Lafferty, D. J. R. (2014). Brown bears and wolves scavenge humpback whale carcass in Alaska. Ursus 25: 8–13. MacNulty, D., Varley, N. & Smith, D. (2001). Grizzly bear, Ursus arctos, usurps bison calf, Bison bison, captured by wolves, Canis lupus, in Yellowstone National Park, Wyoming. Canadian Journal of Zoology 115: 495–498. Mattisson, J., Persson, J., Andrén, H. & Segerström, P. (2011). Temporal and spatial interactions between an obligate predator, the Eurasian lynx (Lynx lynx), and a facultative scavenger, the wolverine (Gulo gulo). Canadian Journal of Zoology 89: 79–89. Milleret, C. (2011). Estimating wolves (Canis lupus) and brown bear (Ursus arctos) interactions in Central Sweden. Does the emergence of brown bears affect wolf predation patterns? Masters thesis, Université Joseph Fourier–Grenoble. Milleret, C., Ordiz, A., Chapron, G., et al. (2018). Habitat segregation between brown bears and gray wolves in a humandominated landscape. Ecology and Evolution 8: 1–17. Monteith, K. L., Bleich, V. C., Stephenson, T. R. & Pierce, B. M. (2014). Life-history characteristics of mule deer: effects of nutrition in a variable environment. Wildlife Monographs 186: 1–64. Mosolov, V. & Valentsev, A. (2003). Lynx of Eurasia. The Kamchatka. In: The lynx. Regional features of ecology, use and protection (pp. 408–423). Moscow: Nauka. Murphy, K., Felzien, G., Hornocker, M. & Ruth, T. (1998). Encounter competition between bears and cougars: some ecological implications. Ursus 10: 55–60. Ordiz, A., Bischof, R. & Swenson, J. E. (2013). Saving large carnivores, but losing the apex predator? Biological Conservation 168: 128–133.

43

Systematics, Ecology, and Behavior

Ordiz, A., Milleret, C., Kindberg, J., et al. (2015). Wolves, people, and brown bears influence the expansion of the recolonizing wolf population in Scandinavia. Ecosphere 6: 1–14. Ordiz, A., Sæbø, S., Kindberg, J., Swenson, J. E. & Støen, O. G. (2017). Seasonality and human disturbance alter brown bear activity patterns: implications for circumpolar carnivore conservation? Animal Conservation 20: 51–60. Palomares, F. & Caro, T. M. (1999). Interspecific killing among mammalian carnivores. The American Naturalist 153: 492–508. Périquet, S., Fritz, H. & Revilla, E. (2015). The lion king and the hyaena queen: large carnivore interactions and coexistence. Biological Reviews 90: 1197–1214. Pikunov, D. G. & Korkishko, V. G. (1992). Leopard of Far East. Moscow: Nauka [in Russian]. Rakov, N. V. (1965). The current distribution of the tiger in the AmurUssuri region. Zoologicheskii Zhurnal 44: 433–441. Ruong, I. (1982). Samerna i historien och nutiden. Stockholm: BonnierFakta. Sand, H., Wabakken, P., Zimmermann, B., et al. (2008). Summer kill rates and predation pattern in a wolf–moose system: can we rely on winter estimates? Oecologia 156: 53–64.

Seryodkin, I. V., Miquelle, D. G., Goodrich, J. M., Kostyria, A. V. & Petrunenko, Y. K. (2018). Interspecific relationships between the Amur tiger (Panthera tigris altaica) and brown (Ursus arctos) and Asiatic black (Ursus thibetanus) bears. Biology Bulletin 45: 853–864. Sinclair, A. R. E. & Krebs, C. J. (2002). Complex numerical responses to top-down and bottom-up processes in vertebrate populations. Philosophical Transactions of the Royal Society of London B 357: 1221–1231. Sivertsen, T. R. (2017). Risk of brown bear predation on semi-domesticated reindeer calves interactions and landscape heterogeneity. Doctoral thesis. Swedish University of Agricultural Sciences, Uppsala, Sweden. Smith, D. W., Peterson, R. O. & Houston, D. B. (2003). Yellowstone after wolves. BioScience 53: 330–340. Stirling, I. & Derocher, A. (1990). Factors affecting the evolution and behavioral ecology of the modern bears. Bears: Their Biology and Management 8: 189–204.

Tallian, A., Ordiz, A., Metz, M. C., et al. (2017). Competition between apex predators? Brown bears decrease wolf kill rate on two continents. Proceedings of the Royal Society B: Biological Sciences 284: 20162368. Tkachenko, K. N. (2012). Specific features of feeding of the Amur tiger Panthera tigris altaica (Carnivora, Felidae) in a densely populated locality (with reference to Bol’shekhekhtsirskii Reserve and its environs). Biology Bulletin 39: 279–287. Wikenros, C., Sand, H., Ahlqvist, P. & Liberg, O. (2013). Biomass flow and scavengers use of carcasses after recolonization of an apex predator. PLoS One 8: e77373. Wilmers, C. C. & Getz, W. M. (2005). Gray wolves as climate change buffers in Yellowstone. PLoS Biology 3: e92. Wilmers, C. C., Crabtree, R. L., Smith, D. W., Murphy, K. M. & Wayne, M. (2003). Trophic facilitation by introduced top predators: grey wolf subsidies to scavengers in Yellowstone National Park. Journal of Animal Ecology 72: 909–916. Wisz, M. S., Pottier, J., Kissling, W. D., et al. (2013). The role of biotic interactions in shaping distributions and realised assemblages of species: implications for species distribution modelling. Biological Reviews 88: 15–30. Wyman, C. (2002). Grizzly on a bull bison in Yellowstone National Park. Ursus 13: 375–377.

Sanz-Pérez, A., Ordiz, A., Sand, H., et al. (2018). No place like home? A test of the natal habitat-biased dispersal hypothesis in Scandinavian wolves. Royal Society Open Science 5: 181379.

Swenson, J. E., Wabakken, P., Sandegren, F., et al. (1995). The near extinction and recovery of brown bears in Scandinavia in relation to the bear management policies of Norway and Sweden. Wildlife Biology 1: 11–25.

Sato, Y., Mano, T. & Takatsuki, S. (2005). Stomach contents of brown bears (Ursus arctos) in Hokkaido, Japan. Wildlife Biology 11: 133–144.

Swenson, J., Dahle, B., Busk, H., et al. (2007). Predation on moose calves by European brown bears. Journal of Wildlife Management 71: 1993–1997.

Schoener, T. W. (1983). Field experiments on interspecific competition. The American Naturalist 122: 240–285.

Zager, P. & Beecham, J. (2006). The role of American black bears and brown bears as predators on ungulates in North America. Ursus 17: 95–108.

Swenson, J. E., Schneider, M., Zedrosser, A., et al. (2017). Challenges of managing a European brown bear population; lessons from Sweden, 1943–2013. Wildlife Biology 1: wlb.00251.

Zhiryakov, V. A. & Baidavletov, R. Z. (2003). Lynx of Eurasia. Kazakhstan. In: The lynx. Regional features of ecology, use and protection (pp. 344–367). Moscow: Nauka.

Schwartz, C. C., Cain, S. L., Podruzny, S., Cherry, S. & Frattaroli, L. (2010).

44

Contrasting activity patterns of sympatric and allopatric black and grizzly bears. Journal of Wildlife Management 74: 1628–1638.

Zabel, A. & Holm-Müller, K. H. (2008). Conservation performance payments for carnivore conservation in Sweden. Conservation Biology 22: 247–251.

Chapter

4

Adaptations and Competitive Interactions of Tropical Asian Bear Species Define Their Biogeography: Past, Present, and Future Robert Steinmetz, David L. Garshelis, and Anwaruddin Choudhury

Three potentially competing bear species inhabit tropical Asia: the sloth bear (Melursus ursinus), sun bear (Helarctos malayanus), and Asiatic black bear (henceforth black bear; Ursus thibetanus) (Figure 4.1). Sun bears (30–80 kg), the smallest species of bear in the world, are about half the size of black bears (65–150 kg) and sloth bears (55–145 kg). Sloth bears occur in the Indian subcontinent (Nepal, India) and Sri Lanka. Black bears range from the Russian Far East to South-East Asia and west to Iran, crossing northern and north-eastern India. Sun bears occupy mainland South-East Asia and the two large islands of Sumatra and Borneo, but only extend west to northeastern India and north to Yunnan, China. Sloth bears and sun bears are wholly tropical–subtropical species, whereas black bears range widely into temperate regions of Asia. Geographic ranges of the black bear and sun bear overlap extensively. Ranges of the sloth bear and black bear abut, and slightly overlap, along the edges of their respective ranges in the Terai arc of southern Nepal and northern India, and may have once overlapped more in north-eastern India. The ranges of all three species converge in north-east India (Figure 4.2). What factors generate the separation of sloth bears geographically from black and sun bears? What factors facilitate the extensive sympatry of black bears and sun bears? How are these patterns structured by evolutionary history and competition between bear species, and what mechanisms facilitate their coexistence or maintain their separation? Has current forest loss and degradation benefited one species over another? If so, has interspecific competition played a part? These questions are the focus of this chapter. The spatial distribution of a species is determined by three factors: (1) evolutionary history and dispersal capacities; (2) the distribution of resources and environmental conditions favorable to survival and reproduction; and (3) the biotic environment involving predators, pathogens, and competitors (Wiens 2011; Louthan et al. 2015). The necessary resources and conditions (factor 2) constitute the fundamental niche of a species. Species distributions can be constrained by competitors (factor 3), causing realized niches to be smaller than fundamental niches (Holt 2009). In terms of biogeography this means that places outside the observed distribution of a species might be suitable (in terms of resources, climate, etc.), but are occupied by a competitor that is better able to use and

monopolize the resources, or physically exclude occupation by another similar species.

Feeding Habits of Tropical Asian Bears All three tropical bear species are omnivorous, yet there are distinct differences in their diets. Each consumes both fruits and insects, although in different proportions. The sloth bear possesses a unique suite of morphological adaptations specifically for myrmecophagy (ant- and termite-eating): the absence of two upper incisors, a long snout with a vaulted palate, mobile protrusible lips, and closable nostrils (Garshelis et al. 1999; Sacco & Van Valkenburgh 2004). These features enable sloth bears to efficiently suck in large quantities of termites and ants (see also Chapter 9). This capability is so impressive (and social insects abundant in its habitat; Figure 4.3) that the sloth bear is able to subsist almost solely on a diet of insects, violating an otherwise general pattern in carnivore energetics that constrains invertebrate-feeding species to body masses less than 22 kg (Carbone et al. 1999). Fruits, when abundant, may comprise up to 90% of the sloth bear diet; however, termites, ants, and other insects comprise >80% of the diet during non-fruiting periods (Seidensticker et al. 2011), and may dominate the diet year-round in some areas (Joshi et al. 1997). The sun bear too possesses morphological features facilitating insectivory (see also Chapter 8). Proportionate to its body size, the sun bear has especially large, strong canines (Christiansen 2008). The skull is wide, supporting large masseter muscles, giving the sun bear an exceptionally powerful bite force. Together, these features might be functional adaptations for breaking into hard termite colonies and for tearing into the trunks of tropical hardwood trees to access nests of stingless bees (Fredriksson 2012). Alternately, these adaptations might have evolved primarily for self-defense, as the sun bear is small-bodied and must contend with potentially dangerous predators including tigers, leopards, dholes, and black bears; nevertheless, their teeth and skull certainly function well to enhance insect-feeding. When fruit is scarce, as occurs periodically between masting events in Indonesia and Malaysia, insects are the dietary mainstay for sun bears; but when fruits are sufficiently available, sun bears are almost

45

Systematics, Ecology, and Behavior

Figure 4.1 Global distributions of Asiatic black bears, sloth bears, and sun bears. Maps are from the Bear Specialist Group of IUCN (International Union for Conservation of Nature), combining definite and possible range.

completely frugivorous (Wong et al. 2002; Fredriksson et al. 2006; Steinmetz et al. 2011). Black bears feed mainly on fruit (Steinmetz et al. 2011). Insects generally constitute a small proportion of their diet (0–4% relative volume) throughout their range (Hwang et al. 2002; Huygens et al. 2003; Garshelis, 2004; see also Chapter 10). Thus, both sun bears and black bears feed predominantly on fruit when available, with sun bears in particular falling back on insects when fruit is scarce (Hwang et al. 2002; Wong et al. 2002; Fredriksson et al. 2006). A key difference between these species is their ability to subsist for an extended time on insects. Sloth bears clearly can. Although sun bears are highly insectivorous when necessary, they become emaciated and some even starve during lengthy fruiting failures (Wong et al. 2005; Fredriksson 2012). Black bears also regularly eat insects during seasons with low fruit availability (Koike 2010), but they cannot meet their energy demands on an exclusively insect diet (Yamazaki et al. 2012).

46

Thus, black bears and sun bears depend on fruits, whereas sloth bears can efficiently exploit fruit-poor habitats with an abundance of insects. These characteristics of morphology, feeding habits, and niches have largely dictated the geographical distributions and ecological overlap of the three species.

Range Limits, Overlap, and Separation The sloth bear is restricted to the Indian subcontinent, with at least 90% of its range in India. There is no evidence, presently or historically, that this species ranged east into Myanmar. The black bear and sun bear both occur in Myanmar and across South-East Asia, but the black bear does not range south into Malaysia, where rainforests replace seasonal tropical forests. Sun bears continue south into the Sundaic region, which includes the southern Malay Peninsula and the islands of Sumatra and Borneo, but northward they do not cross the Red River in northern Vietnam or the Brahmaputra in Assam, India.

Interspecific Competition

(a)

Figure 4.2 Distribution and overlap of Asiatic black bears, sloth bears, and sun bears in north-east India, (a) historically, ca. 1950, and (b) currently. This region of convergence represents the eastern extent of sloth bears and western extent of sun bears, and the only place where all three tropical Asian bears once overlapped. The data are gleaned from various sources of literature, interviews, photographs, and specimens, and point locations were connected to form estimated ranges.

Map by Anwaruddin Choudhury

India

Brahmaputra River

N Ursus thibetanus Melursus ursinus Helarctos malayanus

0

100 km

(b)

Map by Anwaruddin Choudhury

India

Brahmaputra River

N Ursus thibetanus Melursus ursinus Helarctos malayanus

0

100 km

Sloth bear and black bear distributions overlap in one small area in the Indian state of Uttarakhand (including Corbett and Rajaji National Parks) within the Terai arc, although on a fine scale the actual overlap may be quite limited (Seidensticker et al. 2011). In Nepal, the ranges of these two species are mainly

allopatric, with black bears in the Middle Hills and sloth bears in the lowland Terai, but recently a few black bears have been camera-trapped in two Terai national parks, Bardia and Banke (Yadav et al. 2017). Moving eastward, their ranges diverge such that only sloth bears occur in Chitwan National Park, Nepal,

47

Systematics, Ecology, and Behavior

Density of termite mounds per ha

30 25

Sloth bear only Sun bear only

20 15

Fossils Shed Light on Past Ranges

10 5

Sun + Black bear

0 Indonesia Thailand (Kalimantan) (Thung Yai)

Laos (central)

India (Madya Nepal Pradesh) (Chitwan)

Figure 4.3 Differences in availability of epigeal (mound-building) termites in different geographic zones inhabited by the tropical Asian bears. Termite data are from: Indonesia (Mathisen 2003), Thailand (Steinmetz et al. 2011), Laos (Miyagawa et al. 2011), central India (Akhtar et al. 2004), and Nepal (Joshi et al. 1997).

and only black bears occur in Bhutan. Just south of Bhutan, in Manas National Park, India, the two species co-occur. Within north-eastern India there is a patchwork (Figure 4.2b), with some areas occupied by only one species (e.g. only sloth bears in Kaziranga National Park, with one record of a black bear), and some occupied by two species (black bears with either sloth bears or sun bears). Historically, there appears to be a limited degree of overlap between sloth bears and sun bears (Figure 4.2a). However, the distribution of each of these species in north-east India remains somewhat uncertain because they are often confused, by local people, by government authorities, and even in camera trap photos. This makes many reports of species overlap suspect, both presently and historically. It is unclear whether there is (or was) any area where all three species share(d) the same portion of forest. There is tentative evidence that all three species may occur in northern parts of Karbi Anglong in Assam, during the rainy season when sloth bears purportedly move there from flooded areas in Kaziranga (Choudhury 2001, 2011). The distributions of black bears and sun bears in mainland South-East Asia (Myanmar, Thailand, Laos, Cambodia, and Vietnam) are better known. Within this area, which spans millions of square kilometers, sun bears and black bears commonly co-occur within most habitats. Black bear range mainly coincides with broadleaved forests. They are limited northward (in the Russian Far East; Figure 4.1) by boreal coniferous forest (occupied by brown bears, Ursus arctos), and limited elevationally by treeline. They occur up to 4300 m in India, and possibly even higher in Nepal. In mainland South-East Asia, black bears are commonly found in montane habitats; conversely, sun bears are more commonly found at lower elevations but tend to be rare at high elevations (Steinmetz 2011; Steinmetz et al. 2011). However, in Borneo and Sumatra, where black bears are absent, sun bears reach to at least 2000 m (Tumbelaka & Fredriksson 2006); they were frequently camera-trapped at

48

700–1940 m in Sumatra (Linkie et al. 2007). The sloth bear is primarily a lowland species, occurring mostly below 500 m. This elevation range corresponds to termite and ant availability, which declines sharply with increasing elevation in the tropics (Collins 1980).

Are the geographic ranges of tropical Asian bears recent arrangements, or have they been this way for millennia? Sloth bears, black bears, and sun bears originated from a common ancestor in Eurasia during the Miocene to Pliocene transition, 5–6 million years ago (Krause et al. 2008). This period was characterized by the global climate becoming cooler and drier, which caused the fragmentation of forests and expansion of savanna and grassland (Zachos et al. 2001; Woodburne 2004). Savanna ecosystems are often rich in insect abundance but poor in fruit abundance; they can support extremely high densities of social insects like termites (Ferrar 1982). A food this small yet so abundant would be an evolutionary pressure promoting feeding specializations for rapid consumption, and other adaptations for living on a diet low in carbohydrates. The evolution of ecological specialization tends to be favored by the abundance (as opposed to scarcity) of a particular environmental state or resource (Futuyma 1988). It may be that this new niche, afforded by a habitat that was rich in insect availability but likely poor in fruit abundance, promoted development of the unique feeding adaptations of the sloth bear. These specialized feeding habits may have also limited the sloth bear to the region where the food to which they were adapted was sufficient to sustain them. Few sloth bear fossils have been found, and, with one possible exception, all occur within the current range of this species. A fossil bear named Ursus theobaldi, identified only from a few damaged teeth, might actually be a sloth bear. Its location in north-western India (Erdbrink 1953), well beyond the current distribution, indicates their distribution might have been wider historically, and perhaps overlapped more extensively with black bears, which currently inhabit that area and presumably did so in the past. Whereas sloth bears specialized, and accordingly thrived only in the region where termites and ants were abundant, black bears remained generalists, little evolved from the prototypic form. Fossil remains of this species are difficult to distinguish from Ursus minimus, their presumed ancestor, and have led to many apparent misinterpretations of their former range (Wagner et al. 2011). Purported U. thibetanus fossils have been found across a very broad area, as far north as the Ural Mountains (Russia) and Germany, and as far west as France, dating from the early Pliocene to late Pleistocene; however, it now seems likely that this species originally evolved in Asia and spread to Europe for a time in the early Pleistocene. Black bear fossils from the Middle and Late Pleistocene are widespread in Asia, occurring in Vietnam, China, and Thailand (Tougard 2001), but do not extend south into Malaysia or Indonesia, or into the Indian subcontinent.

Interspecific Competition

In contrast to the black bear, sun bear fossils are scarce in mainland Asia. However, there are three locales where both sun bear and black bear fossils have been found: two in northern Vietnam (Tougard 2001; Bacon et al. 2008), and one in Shanxi province, northern China (well north of the presumed Pleistocene range for this species, creating some doubt as to the location of this find; Erdbrink 1953). Late Pleistocene sun bear fossils are far more widespread in Sumatra, Borneo, and Java (where they went extinct less than 10,000 years ago) (Bacon et al. 2008). This fossil history imparts the following conclusions relevant to species interactions. First, there is no evidence that ranges of black bears and sloth bears ever broadly overlapped. That is, their current separation, with marginal overlap in places, does not seem to result from range contraction, but rather sloth bears evolving to take advantage of an area where black bears could not live. Second, black bears and sun bears have coexisted in mainland South-East Asia for hundreds of thousands of years. They have similar ecological needs, yet one does not outcompete and exclude the other, at least on a broad scale. Third, sun bears never ranged west beyond north-east India (claims of their occurrence in Nepal were in error). The western ecological limits for this species remain unclear, but suffice to say that north-east India is the western limit of sun bears, the eastern limit of sloth bears, and central within the range of black bears, making it an intriguing area of interest (Figure 4.2).

Why are Sloth Bears Disjunct From Black Bears and Sun Bears? The black bear occupies a wide range of habitats, from an arid, sparsely vegetated landscape in Iran to tropical forests of SouthEast Asia to high elevations in the Himalayas. Likewise, the sun bear occupies habitats as diverse as dry forest with sparse tree cover in Laos, to dense rainforest with towering canopies of dipterocarps in Borneo. It would seem that the fundamental niches of black bears and possibly sun bears would include forested habitats of peninsular India, and that they have had sufficient time historically to reach the area – so why don’t they occur there? We believe there are two reasons: (1) low abundance and diversity of fruits, and (2) competitive exclusion by sloth bears. Sloth bear habitats in the Indian subcontinent appear to be particularly poor in fruit availability, yet relatively rich in termites (Figure 4.3), a food to which the species is uniquely adapted. For example, average density of fruit trees in a sloth bear-occupied sal forest in India was 38 per ha, or just 19% of overall tree density (Akhtar et al. 2004), whereas at a site in Thailand (where sloth bears are absent but black and sun bears coexist) there were 158–223 fruit trees per ha, comprising 40–53% of overall tree density (Steinmetz et al. 2013). In contrast, termites are 6–20 times more abundant in sloth bear habitat than outside sloth bear range (Figure 4.3). Over evolutionary time scales, the prior adaptation and occupation by one species can preclude another species from evolving to occupy the same habitat (Wiens 2011). This process, known as niche pre-emption, integrates abiotic and biotic

influences on species distribution (Wiens 2011). When the black bear arrived at the boundary of sloth bear-occupied territory, it was excluded from invading this region by the existing occupation of a more specialized species that was well-adapted for efficient exploitation of the resources there (Futuyma 1988). Unlike in mainland South-East Asia, where abundant fruit, combined with the sun bear’s ability to supplement its diet with less-preferred insects (see below) facilitates coexistence of black bears and sun bears, sloth bear range is mainly rich in insects, and fruit abundance is too scanty to support another frugivorous bear species (Figure 4.4). Another example of niche pre-emption occurred in Japan. Black bears dispersed into Japan from continental Asia during the mid-Pleistocene, but remained within a small region of Honshu Island, even after tens of thousands of years, seemingly constrained by the larger-bodied brown bear, which already occupied much of the region (Wu et al. 2015). Fossil evidence and mtDNA indicate that the brown bear went extinct on Honshu Island during the late Pleistocene, after which the black bear expanded its range (but never occupied Hokkaido Island, where brown bears still exist). The extinction of the brown bear, probably a major potential competitor, on Honshu apparently opened niche space for the black bear, which until then had been excluded from otherwise suitable habitat. Niche pre-emption does not require the winning species to be larger. For example, among the many islands off the coast of Alaska, some are occupied by brown bears and some (mainly smaller ones) by American black bears (U. americanus), and there is evidence that very high densities of the latter, smaller species preclude colonization by brown bears (Mattson et al. 2005). In India, we posit that whereas sloth bears and black bears are similar in size, sloth bears gained an advantage by colonizing the subcontinent earlier, and likely reached high densities there, as they require a relatively small amount of space (Garshelis et al. 1999). They also maintained a competitive advantage in their ability to thrive in this fruit-poor, insect-rich environment. Both of these advantages precluded occupancy by black bears (and possibly sun bears). We now turn to the opposite question: what limits sloth bears to the north (where black bears occur) and east (where both black bears and sun bears occur)? The answer might again be related to food supply in relation to competitive advantages. The uplands and mountains to the north and east of sloth bear range appear to have much lower termite abundance than the Indian subcontinent (Garshelis et al. 2015; Figure 4.3). Thus, the sloth bear’s advantage in termite-rich regions wanes outside of its current range, and this is perhaps exacerbated further by competition from black bears and sun bears.

Why do Black Bears and Sun Bears Broadly Overlap? Studies in Thailand have shown that black bears and sun bears use the same habitats and both rely mostly on fruit, tracking the same fruiting tree species through time (Ngoprasert et al. 2011;

49

Systematics, Ecology, and Behavior

Black bear

Too cold for sun bear (outside the niche of sun bear)

Sloth bear Niche pre-emption leading to competitive exclusion of black bear and sun bear

Shared preference niche allowing coexistence of sun bear and black bear

Sun bear Fruit scarce, termites abundant

Steinmetz et al. 2013). However, some important differences were also found: sun bears were the main consumer of insects, and black bears predominated in fruit-rich montane forest (Steinmetz et al. 2011, 2013). Yet in the Sundaic region, where black bears are absent, sun bears are relatively common in montane forest, suggesting that their low density in montane habitats on the mainland may be due to competitive exclusion by the larger black bears. Sun bears, because of their smaller size and lower absolute food requirements, may be better able than black bears to subsist on insects as a fallback food. In north-east India, sun bears consume an almost equal proportion of insects and plants year-round (Sethy & Chauhan 2018). Niche partitioning can arise under such circumstances, whereby species share preferences for a resource or habitat, but differ in their tolerances or competitive abilities, so become differentially distributed along an environmental gradient (Rosenzweig 1991). The coexistence of black bears and sun bears appears to be structured this way, with overlapping fundamental niches combined with asymmetric competition, known as the “shared preference niche” (Rosenzweig 1991). Extensive niche overlap between these two species may also reflect an abundance of food: they share resources without really competing (Wiens 1993). Black bears may be limited from occupying the Sundaic region not because of sun bears, but because of fruiting cycles (Figure 4.4). Black bears reach their southernmost limit at about 8°520 N in Thailand. This location, 200 km south of the Isthmus of Kra, marks a phytogeographical ecotone between seasonal tropical forests to the north and aseasonal rainforests to the south (Woodruff 2003). South of this point most tree species fruit synchronously during masting events, every 3–7 years, followed by extended intermast periods with very little fruit (Primack & Corlett 2005). Sun bears survive these cycles by periodically switching to insects. In contrast, the seasonal climate of mainland SouthEast Asia, with wet seasons and dry seasons, promotes

50

Fruit abundant and available all year

Figure 4.4 Model of geographic overlap of tropical Asian bear species, as a function of evolutionary history, ecological niches, and resource abundance. Size of circles is roughly proportionate to global range size. Overlap among circles depicts amount of geographic overlap: Asiatic black bears and sun bears overlap extensively, whereas sloth bears overlap little with the others. White boxes explain major mechanisms of ecological separation or overlap. Gray boxes describe niche conditions in different regions of overlap or non-overlap.

Termites abundant, but fruit too low to sustain black bear (outside the niche of black bear)

asynchronous fruiting among tree species, yielding seasonally changing fruit supplies, but no extended periods without fruit. For example, in a primary dipterocarp forest in Indonesian Borneo, an average of 13 trees/ha provided fruit for sun bears during mast fruiting months, whereas only 0.6 trees/ha provided fruit during an intermasting 2-year span (Fredriksson et al. 2006). In contrast, in evergreen and deciduous forests in Thailand, where sun bears and black bears coexist, density of trees with fruit exceeded 10 trees/ha in the rainy season and 6 trees/ha even in the relatively fruitpoor dry season (Steinmetz et al. 2013).

Species Interactions in a Changing World Major differences in the abundance and relative availability of key bear foods across Asia, combined with differences in the species’ evolutionary histories and ecological niches, have produced the biogeographic patterns of distribution of black bears, sun bears, and sloth bears that we see today (Figure 4.4). However, each of these species is under threat from combinations of habitat loss or degradation, poaching for parts, or conflict-related killing. These species may have different vulnerabilities to these threats. Black bears are most sought by poachers for their body parts (gall bladder and paws), but also seem most tolerant of habitat alteration. The south Asian region where the three species intersect is near the center of the geographic range of black bears, but the fringe of the ranges of sloth and sun bears. It is therefore unsurprising that along these fringes, the latter two species have been most vulnerable to habitat change. For example, Yunnan Province in southern China has suffered heavy loss of low and mid-elevational rainforests, which have been converted into plantations of banana, rubber, and tea; sun bears once ranged across a significant portion of this province, but today there is just a single confirmed record of a sun bear in China, 90%.

Development Average neonate birth weight is 112 g. Individual cubs average only 0.12% of the dam’s body weight, the smallest size proportion of any other eutherian mammal (Zhu et al. 2001). Neonates have sparse short, white hair covering the entire body. This altriciality mandates extraordinary maternal investment in care-giving behaviors, including holding the vulnerable neonate continuously for the first 2–3 weeks of life to prevent hypothermia (Zhu et al. 2001; Snyder et al. 2003, 2016). This also explains the dam’s post-partum fast and the cub’s frequent vocalizations during the same time period, probably seeking nipple contact. As in other bears, there is relatively slow maturation of the young (Zhu et al. 2001). The neonate’s skin in the area where black hair will eventually grow turns gray by 8–10 days of age. The typical black and white marking of the adult are evident by 23–25 days when cubs are gaining about 60 g/day. Eyes begin opening from 40 to 49 days and are open by 75 days. The denning interval in the wild is ~4 months, with emergence in early winter. By 6 months of age, cubs gain 110 g daily and are beginning to transition to eating bamboo leaves. The weaning norm in the wild is 18–24 months.

Behavior Due to the altriciality of panda cubs, maternal behavior is notably active, and females engage in a range of behaviors that promote cub suckling, warmth, and comfort. Maternal behaviors are often initiated in response to the plaintive vocalizations cubs emit; however, maternal responsiveness declines over time (Snyder et al. 2003). Birth and initial cub development occur in dens, typically inside tree cavities or in rock caves (Zhu et al. 2001; Zhang et al. 2007; W. Wei et al. 2019). Pandas do not hibernate, and these dens are only used for rearing cubs. Maternal females are highly selective in choosing den sites, preferring cavities of a certain size and shape that protect cubs from the elements and predators. Preferred den sites have a higher capacity to buffer against extremes of temperature and humidity outside, and provide stable microclimatic conditions for rearing vulnerable cubs. Foraging behavior is of the utmost importance to allow pandas to subsist on their plentiful, but low-nutrient diet of bamboo (see Feeding Ecology). The selection, processing, and consumption of bamboo consists of a fairly stereotyped repertoire of behaviors that serve to efficiently exploit bamboo as a

70

food resource (Swaisgood et al. 2018). Most aspects of these specialized behaviors have been found to be consistent between wild-born and captive-born giant pandas, suggesting that maternal mentorship may not drive the acquisition of these skills. While the complex motor patterns associated with efficient processing and consumption of bamboo appear developmentally stable in different rearing environments, pandas’ ability to select the most nutritious parts of the plant (see Feeding Ecology) appears to be compromised during development in captivity. Communication and reproduction are the best-studied aspects of panda behavior. As a solitary-living animal, olfactory (or chemical) communication is the primary mode of communication throughout most of the year (Schaller et al. 1985; Swaisgood et al. 2004). Pandas deposit scent in the form of urine and specialized anogenital gland secretions on trees at traditional sites that serve as a sort of community bulletin board. Pandas are highly selective of their scent deposition sites, preferring some types of trees over others, selecting relatively open microhabitat around scent trees, and selecting certain bark characteristics based on roughness of texture (Nie et al. 2012a). These selective behaviors serve to increase signal range, increase signal duration, and maximize the probability that the signal will be detected by conspecifics. Pandas also deploy four distinct postures to deposit scents, including a handstand posture used by adult males to convey competitive dominance (White et al. 2002). A series of investigations on captive pandas have demonstrated that pandas can extract a great deal of information from investigating scent (Swaisgood et al. 2004). Pandas can discriminate conspecific odors to determine gender, reproductive condition, age class, kin, individual identity, and other important social information. Semiochemical investigations using gas chromatography have shown that panda chemical signals are composed of a complex mixture of chemical constituents that differ between the sexes and change from the breeding to non-breeding seasons (Yuan et al. 2004; Zhou et al. 2019). Acoustic communication, by contrast, is primarily utilized during the mating season (and between mother and cub), and wild pandas vocalize infrequently outside of reproductive contexts (vocalization outside the breeding season is more common in captivity, where pandas are housed near conspecifics and also direct vocalizations towards their caretakers). The complexity of information conveyed in vocal signals, however, is no less pronounced than it is for chemical signals (Charlton et al. 2009, 2010, 2011, 2018a). Panda vocalizations carry information regarding identity, male testosterone levels, and female reproduction condition, and play an instrumental role in courtship and copulation. Among their vocal repertoire, pandas emit high-pitched bleats and chirps, which are closely tied to sexual and affiliative motivation, and lowerpitched moans and barks, which are associated with aggressive intent or perhaps ambivalence between attraction and avoidance. Panda audition (hearing) can be described as typical of large carnivores (Owen, Keating, & Denes 2016). However, giant panda hearing sensitivity diverges significantly from that

Giant Panda

of the polar bear in the upper frequencies, with functional hearing retained into the ultrasonic range. This functionality may support effective intraspecific acoustic communication in the dense bamboo forests that giant pandas inhabit, and may be especially relevant to the transmission of the bleat vocalization through this habitat (Charlton et al. 2018b). Visual communication is less prominent and less wellstudied in pandas. However, some behaviors associated with estrus and leading up to mating are visual signals, such as backward walking, pink vulva, and tail up in females (Owen, Swaisgood, & Blumstein 2016) and footscraping in males. The black and white pelage markings are a prominent feature of pandas, and are thought to play a role in visual communication. While the body markings are likely involved in camouflage, the facial markings may convey information about identity or aggressive intent (Caro et al. 2017). Panda vision is not well studied, but it seems reasonably acute and they possess color vision typical for carnivores (Keller et al. 2006). Presumably, color vision aids in selection of nutritious bamboo, but could also be used in social communication, such as assessment of female reproductive condition via vulva color. Pandas display a large variety of specialized behaviors associated with reproduction (Owen, Swaisgood, & Blumstein 2016) (see Reproduction). These behaviors aid in localizing potential mates, reducing aggression, overcoming a tendency to avoid conspecifics, and supporting the courtship process (Swaisgood et al. 2006); they also allow managers of captive pandas to evaluate the estrous status of females and reproductive readiness. In the wild, when a female comes into estrus she disperses her scent widely, which serves to attract several male suitors that compete for reproductive access (Schaller et al. 1985; Nie et al. 2012b). This intermale competition can last several days, but typically dominance is established in the first 2 days and escalated and injurious aggression is replaced with the use of visual and acoustic signals, and conflict is mediated by threat and avoidance. The largest male typically prevails and only a single male mates with the female. In captivity, both male and female choice have been shown to be important (Martin-Wintle et al. 2015). Much higher rates of copulation and fertilization are achieved when partners are allowed to choose their own mates than when they are paired with randomly selected partners that are less preferred. Personality also plays a role in these captive parings, with certain combinations of partners having better reproductive performance than other personality pairings (Martin-Wintle et al. 2017). These findings are extremely important for guiding reproductive management in conservation breeding programs, but their relevance to mating behavior in wild populations remains unknown.

Parasites and Diseases Parasitism is prevalent in wild panda populations and may be a cause for concern (Zhang et al. 2008). Baylisascaris schroederi is the most common and harmful parasite in the panda. Estimated prevalence in captive and wild pandas ranges from

50% to 100% (Zhang et al. 2008). Although the life cycle of B. schroederi remains unknown, ova are passed in the host feces and into the environment, where they can persist in soil for years, from which they can be ingested (Zhang et al. 2008). Visceral larval migrans (VLM) refers to disease caused by aberrant migration of nematode larvae, which can be profoundly inflammatory. Diagnosis has historically relied on identification of adult worms or eggs in the feces of infected individuals. Recent studies have shown that polymerase chain reaction (PCR) can be used to detect B. schroederi eggs with much higher sensitivity and specificity (Wang et al. 2013). Other species of nematode have been identified in pandas, including Ancylostoma ailuropodae, Ogmocotyle sikae, Toxocaris seleactis, and Strongyloides sp., but their clinical significance remains unknown. Over 12 hard tick species have been described infesting pandas, including ticks from the genera Ixodes, Haemaphysalis, and Dermacentor (Cheng et al. 2013). Although it is currently unknown whether these species act as a vector for diseases afflicting pandas, tick infestation alone can cause morbidity, such as dermatitis, or in extreme infestations may result in anemia, weight loss, and death. Giant pandas can also be infested by Chorioptes mites, which have been recently confirmed to be a unique species, Chorioptes panda (Wang et al. 2012). This mite has been isolated from both wild and captive individuals. Acariasis caused by C. panda presents as erythema, crusting, and alopecia (due to pruritus). Of the four protozoal parasites reported in pandas, Toxoplasma gondii is the only one known to have resulted in clinical morbidity and eventual mortality (Ma et al. 2015). The microsporidian pathogen Enterocytozoon bieneusi is an opportunistic enteric parasite that primarily resides in the small intestine of the host, but no clinical signs have been described in pandas.

Status in the Wild The giant panda is considered Vulnerable by the International Union for the Conservation of Nature (IUCN), although it was listed as Endangered until 2016 (Swaisgood et al. 2016). The Chinese government has also recently downlisted the giant panda to Vulnerable using similar criteria (Zhang et al. 2017b). These conservation status assessments incorporate the best available evidence regarding population size and trends, the amount and trends of supporting habitat, and the degree of population fragmentation threatening a species, among other factors. A Vulnerable species under IUCN criteria still faces high risk of extinction in the wild, so the downlisting of the giant panda does not indicate that further conservation efforts are not warranted. On the contrary, ongoing conservation action is required to address the small population size and continuing and emerging threats facing giant pandas (Kang & Li 2016; Swaisgood et al. 2016, 2017; Xu et al. 2017). The giant panda is also protected against international trade by the Convention on International Trade in Endangered Species of Wild Fauna and Flora (CITES), Appendix 1 (the highest level of protection).

71

Species Accounts

Figure 6.6 Plot-level occurrence of human disturbance in the third (gray; 1999–2003) and fourth (black; 2011–2014) Giant Panda National Surveys in Sichuan Province, China. Proportions of human disturbance and biological covariates that differ between the third and fourth national surveys are indicated by *P < 0.05, **P < 0.01. Reprinted with permission of W. Wei et al. (2018b) and Conservation Letters.

The panda’s distribution today (Figure 6.1) represents a small proportion of its historical range. Pandas once occupied much of southern China and were found as far north as Beijing. Range retraction was associated first with a warming climate toward the end of the Pleistocene (Han et al. 2019), but was greatly expedited during recent history due to rapidly expanding human populations, conversion of habitat for agricultural expansion, and, later, industrialization and urbanization (Zhu et al. 2013; Li et al. 2015b). Remnants of panda populations today are found in the Qinling, Minshan, Qionglai, Liangshan, Daxiangling, and Xiaoxiangling mountain ranges. The status of the giant panda is better known than for most bear species – indeed, than most species of wildlife – because the Chinese government invests in an intensive rangewide national survey about once a decade (SFA 2015). Hundreds of workers spend tens of thousands of worker-days conducting transects through all suitable panda habitat. Pandas are notoriously difficult to find in their dense bamboo habitat, so the census relies on signs, primarily feces. Although DNA census techniques have been developed for pandas (Zhan et al. 2006; Wei et al. 2012), they have not been widely applied in the national survey, and population numbers are estimated using fecal remains. This necessarily crude estimate may underestimate population size, as suggested by comparisons with molecular census methods (Zhan et al. 2006), yet still provides useful information on population trends. Examining the four surveys conducted to date, beginning in the 1980s, reveals a trend of rapid decline, followed by stabilization and a gradual increase in population numbers. The most recent survey, completed in 2014, estimated 1864 individual pandas (with cubs 600 km (Durner et al. 2011), although such long-distance swims can result in increased mortality (Monnett & Gleason 2006), and swimming is energetically expensive (Griffen 2018). Some polar bears employ aquatic stalking to approach basking seals by swimming below ice floes. Bears may use multiple short dives to approach seals; the longest recorded dive by a polar bear approaching a basking seal is 3 min 10 s (Stirling & van Meurs 2015). Polar bears often dive to depths of 3–4 m apparently to feed on seaweed (Lone et al. 2018a); maximum recorded dive depth was 13.9 m and most bears dove to >6 m depth.

Parasites, Infections, and Diseases Most disease-related investigations on wild polar bears are serological surveys documenting the occurrence of antibodies against potential pathogens. Some of these infections are associated with disease and mortality, some have an unknown impact, and some may be transmissible to humans. Concurrent exposure to two or more pathogens at the same time, as well as to persistent organic pollutants (POPs) and other compounds that may act as immunomodulators (Dietz et al. 2018a), may impact the outcome of infections. Most reports of disease in polar bears have been documented in captive animals, often with infectious agents that are less relevant for wild bears (Fagre et al. 2015). However, shrinking ice coverage of the Arctic Ocean has opened new trans-Arctic sea routes for tourism and cargo, also increasing the potential for transmission of new diseases to the region.

Parasites A range of parasites, sometimes causing disease, have been documented in captive polar bears (Fagre et al. 2015). Whether wild polar bears are exposed to these parasites is largely unknown, because the parasitic fauna hosted by polar bears have been scarcely studied with the exception of Trichinella sp. and Toxoplasma.

202

Trichinella is a genus of nematodes that have a wide host range infecting polar bears and other Arctic carnivores. A seroprevalence of Trichinella sp., presumably against the Arctic variant, T. nativa, of typically >50% has been reported from polar bears on the coast of Alaska and the Bering Strait region and in Svalbard and the Barents Sea (Fay 1960; Dau & Barrett 1981; Åsbakk et al. 2010). Some disease outbreaks in humans have been associated with undercooked meat from polar bears and walrus, but the parasite is not often found in seals (Thorshaug & Rosted 1956; Tryland 2000; Tryland et al. 2014). A recent review reported 63 cases of human trichinosis linked to consumption of polar bear meat since World War II. However, although the prevalence of Trichinella sp. in polar bears is reported as high, the number of larvae in the muscular tissues is usually low, diminishing the risk for humans (Dupouy-Camet et al. 2017), and suggesting that the parasite has limited impact on the polar bear host. Toxoplasma gondii is a zoonotic intestinal coccidian parasite of domestic and wild felines. Polar bears, and probably any mammal or bird, may serve as intermediate hosts, carrying extra-intestinal tissue cysts in organs such as muscle, liver, lung, and brain. Clinical toxoplasmosis has been diagnosed in a variety of captive and wild marine mammals (Tryland 2000; Tryland et al. 2014; Miller et al. 2018), affecting the central nervous system and causing reproductive failure (Miller et al. 2018). Serological investigations of polar bears from Alaska, Canada, Svalbard and the Barents Sea and Russia have indicated exposure, but clinical toxoplasmosis (i.e. disease) has not been verified (Miller et al. 2018). Exposure to T. gondii was found to be greater for bears that use land during the summer and fall in the Beaufort Sea (Atwood et al. 2017). However, marine mammals, including polar bears, may represent a source of infection to people, as toxoplasmosis and a high prevalence of antibodies against the parasite have been documented in Inuit communities (McDonald et al. 1990; Messier et al. 2009).

Bacterial Infections Polar bears may host a wide range of potential pathogenic bacteria, but there are not many reports of clinical cases or specific investigations. Many marine mammal species are exposed to bacteria of genus Brucella (Tryland et al. 2018). A seroprevalence of 5–14% has been found in polar bears from Svalbard and the Barents Sea (Tryland et al. 2001; Nymo et al. 2013), and a seroprevalence of 5–10% in polar bears in the Beaufort and Chukchi Seas (Rah et al. 2005; O’Hara et al. 2010). However, little is known with certainty about which Brucella species these animals have been exposed to, or if this exposure has health impacts on polar bears. In Phocidae (i.e. “true seals”) no pathology has been associated with the isolation of Brucella bacteria, but in Otariidae (eared seals), B. pinnipedialis has been associated with pathology in reproductive organs (Sidor et al. 2008; Duncan et al. 2014).

Polar Bear

Viral Infections A wide range of viral infections have been documented and may also cause disease in marine mammals (Duignan et al. 2018), but information about such infections and possible impacts in polar bears is scarce. Rabies virus (family Rhabdoviridae, genus Lyssavirus) is neutrotrophic, has a wide global distribution, and can cause rabies in a wide range of host species. One clinical case of rabies in a polar bear is published. The bear was found (Northwest Territories, Canada) by Inuit hunters, dragging its hind legs, and the diagnosis was based upon submission of the spinal cord and head and the finding of specific staining of inclusion bodies in neurons of the lumbar spinal cord (Loewen et al. 1990; Taylor et al. 1991). A serological screening of 266 polar bears in Svalbard and the Barents Sea revealed no seropositive animals (Tryland et al. 2005). In spite of the circumpolar distribution of the rabies virus and the Arctic fox (Vulpes lagopus) being its main reservoir, it is assumed polar bear exposure to rabies is low. Morbillivirus (family Paramyxoviridae, genus Morbillivirus) has caused severe disease and epizootics in several seal and some whale species, and serological investigations have revealed exposure of many species and populations (Duignan et al. 2018). Antibodies against a variety of morbilliviruses (canine distemper, CDV; phocine distemper, PDV; dolphin morbillivirus, DMV; porpoise morbillivirus, PMV) have been detected in wild polar bears (3–50%) in Alaska, Canada, Svalbard/Barents Sea, and Russia (Fagre et al. 2015). As these viruses have similar antigenic structures and to some extent cross-react serologically, the tests have often been conducted with little evidence of which virus animals were exposed to (Fagre et al. 2015). There are indications that morbillivirus infections in polar bears may cause immunosuppression (Kirk et al. 2010), and thus may impact the course of infection with other pathogens. Antibodies against calicivirus have been reported in polar bears from Canada and from Svalbard and the Barents Sea (Philippa et al. 2004; Tryland et al. 2005), but it is not known which type of calicivirus they were exposed to or if it has caused disease. Polar bears from Canada were also exposed to canine adenovirus (CAV-2) or a similar virus (Philippa et al. 2004). In captive polar bears, it has been shown that equine herpesvirus 9 and suid herpesvirus 1 both were able to infect and cause disease, but exposure was probably through other corralled animal species or via the feed (Fagre et al. 2015), and is of little relevance for wild polar bears. A captive polar bear in Toronto, Canada was diagnosed with West Nile virus after presenting with non-ambulatory paraparesis (Dutton et al. 2009).

Industrial Chemicals The polar bear is among the species most contaminated with mercury and persistent organic pollutants (POPs) both in the Arctic and globally due to its position at the top of the marine

food web (Letcher et al. 2010; Dietz et al. 2013, 2018; Bignert et al. 2016; de Wit et al. 2017; Routti et al. 2019). Since 2001, numerous POPs have been listed, and thus globally banned, under the United Nations Environmental Programme’s (UNEP’s) Stockholm Convention on POPs, as well as mercury under UNEP’s Minamata Convention on mercury. Given their high trophic level, the polar bear is a useful indicator species for biomonitoring spatiotemporal trends, distribution, dynamics, fate, biomagnification, and potential effects of mercury (metals) and legacy and new POPs (Letcher et al. 2010; Dietz et al. 2013, 2018; Rigét et al. 2019; Routti et al. 2019). These POPs originate mainly from industrial and agricultural activities occurring at more southern latitudes, and are transported by air, ocean currents, and river outflows to the Arctic. Although local pollution sources exist in the Arctic, their influence is seen only in a limited geographical area and their contribution to larger-scale contamination is likely minor (Brown et al. 2018). As has been detailed in recent reviews, the majority of contaminant-related studies have focused on contaminant “hot spots” and specifically in the Beaufort/Chukchi Sea, Hudson Bay, East Greenland, and Svalbard regions of the Arctic (Letcher et al. 2010; Bignert et al. 2016; de Wit et al. 2017; Dietz et al. 2013, 2018; Routti et al. 2019). Levels of some legacy POPs such as polychlorinated biphenyls (PCBs) in polar bears have been shown to have decreased prior and up to 2000, but have since remained relatively unchanged and in some cases are shown to be increasing due to changes in their diet and underlying food web (Dietz et al. 2013, 2018, 2019; Bignert et al. 2016; Brown et al. 2018; Letcher et al. 2018). A number of emerging environmental pollutants, such as polybrominated diphenyl ether (PBDE) flame retardants and per-/polyfluoroalkyl substances (PFASs; in particular the highly bioaccumulative perfluorooctane sulfonic acid, PFOS) among others have been reported (de Wit et al. 2017). In a recent study from Hudson Bay subpopulations, 247 POPs were screened in liver or fat samples and 210 POPs were detectable (Letcher et al. 2018). Liu et al. (2018) detected hundreds of analytes in the serum of polar bears from Hudson Bay and the Beaufort Sea, which belong to 13 classes of chemicals, including novel PCB metabolites and many fluorinated or chlorinated substances not previously detected in polar bears. These studies and numerous others (Bignert et al. 2016; de Wit et al. 2017 and references therein) exemplify that polar bears in general are exposed to an increasingly complex array of POPs via multiple routes of exposure but mainly through the diet. Although overall levels are high, concentrations of PCBs and most organochlorines show wide variation between individual polar bears, which may be related to age, sex, breeding status, changes in body fatness and energetic requirements, food web length and diet composition, and finally individual differences in biotransformation capacity (Letcher et al. 2010; de Wit et al. 2017; Routti et al. 2019). Effects of global warming have the potential to result in interactive effects on polar bears resulting from nutritional

203

Species Accounts

stress, novel pathogen exposure, changing bear–human interactions (e.g. increased land use, increased shipping) and exposure to POPs, mercury, and other metals (Letcher et al. 2010; McKinney et al. 2010, 2013, 2015; Bechshøft et al. 2018; Dietz et al. 2018). Biological effects assessments provide evidence regarding the possible effects of POP exposure on physiological pathways in polar bears, including endocrine and reproductive challenges, immunological impairment, and alterations to vitamin homeostasis, among others (Letcher et al. 2010; Sonne 2010; Sonne et al. 2012; Dietz et al. 2018). Effect studies of Arctic biota that have been documented for old POPs, mercury (and a few other metals), flame retardants, and a few PFAS (mostly on PFOS) have been entirely based on correlative relationships between biomarker/effect endpoints and measurements of POPs in body tissues (liver, fat, and muscle), as well as in blood plasma (Letcher et al. 2010; Sonne 2010; Dietz et al. 2018). Thus, the suggestion of effects is based on a weight of correlative evidence rather than cause–effect relationships. Furthermore, consideration of combined or “complex mixture” effects which include targeted and as yet unidentified and new POPs (de Wit et al. 2017) will advance the understanding of the impacts of contaminants as a consequence of using these more environmentally realistic exposure scenarios for polar bears and Arctic wildlife in general.

Status in the Wild Polar bear subpopulations vary in size from a few hundred to a few thousand bears with a total world population estimated at about 26,000 (Regehr et al. 2016). The quality of abundance estimates for individual subpopulations varies greatly from recent rigorous estimates to estimates that are dated, and others which are based on “best guesses.” Because of this, the global population estimate is not used as a monitoring benchmark. Rather, it expresses a reasonable range based on a combination of the best available information and understandings of polar bear habitat. The polar bear was listed as Vulnerable under criterion A3c (a suspected population reduction during the coming three generations based on a decline in the area of occupancy, extent, and habitat quality) by the IUCN in 2015 (Wiig et al. 2015). This designation was based on analysis that estimated 0.71 probability (range 0.20–0.95) of mean global population size decline 30% over three generations (Regehr et al. 2016). Declines in sea-ice habitat have been associated with declines in polar bear body condition, survival, and abundance in some subpopulations (Stirling et al. 1999; Regehr et al. 2007, 2010; Rode et al. 2010, 2014; Bromaghin et al. 2015; Wiig et al. 2015; Lunn et al. 2016; Obbard et al. 2016). In Canada, where 13 of the 19 subpopulations occur and the total estimate is about 15,000 bears, the polar bear was listed as of Special Concern in 2018 (COSEWIC 2008). Three Canadian subpopulations are currently listed as “Likely Declined,” two as “Likely Stable,” one as “Stable,” one as

204

“Likely Increased,” one as “Increased,” and five are listed as “Data Deficient/Uncertain” (Environment and Climate Change Canada 2018). Polar bear harvest is legal in Canada both for subsistence and sport hunting (in some subpopulations). In the United States, the US Fish and Wildlife Service listed the polar bear as “threatened” worldwide under the US Endangered Species Act in 2008 due to range-wide declines in sea ice (US Fish and Wildlife Service 2008). The United States shares management responsibility for the Southern Beaufort Sea subpopulation with Canada (listed by Canada as Likely Declined) and for the Chukchi Sea subpopulation with the Russian Federation. Polar bear harvesting is legal in the US by coastal Native hunters as outlined in the US Marine Mammal Protection Act. The Southern Beaufort Sea subpopulation declined by between 25% and 50% in the period 2001–2010 (Bromaghin et al. 2015). In contrast, a recent study estimated that there were about 2900 bears in the Chukchi Sea subpopulation (Regehr et al. 2018), and that the subpopulation has been productive in recent years. Hunting of polar bears was banned in the former Soviet Union in 1956 (Larsen & Stirling 2009). Currently, the polar bear is listed in the Red Data Book of the Russian Federation (Belikov et al. 2010); thus, the polar bear is given special protection and any action that causes death or decline in the abundance of polar bears is prohibited. A US–Russia Bilateral Treaty was signed between Native and government representatives of the US and Russia in 2000 and implementation began in 2007. The polar bear has been protected in Norway since 1973 (Vongraven et al. 2010); therefore, there is no harvest of the Barents Sea subpopulation in Norwegian territories. In 2006, the polar bear was listed as Vulnerable on the Norwegian Red List due to sea ice loss. The Barents Sea subpopulation was surveyed in 2004 and 2015, and although surveys were not directly comparable due to inadequate access to Russian territory in 2015, the subpopulation is at present considered to be stable (Aars et al. 2017), although the authors suggest that the carrying capacity of the area most likely has been significantly reduced. The Government of Greenland is the responsible authority for the management of polar bears in Greenland, including the national legislation, national coordination and setting of hunting quotas (Jessen 2018). Greenland has sole management responsibility for the East Greenland subpopulation and shared responsibility with Canada for the Kane Basin (KB), Baffin Bay (BB), and Davis Strait (DS) subpopulations. Recent studies have produced abundance estimates for KB and BB (SWG 2016) and DS (Peacock et al. 2013), and to date no abundance estimate is available for the East Greenland subpopulation. A Memorandum of Understanding between Greenland and Canada/Nunavut on the Kane Basin and Baffin Bay subpopulations governs management of these areas, including harvest (Jessen 2018). Greenland listed polar bears as Vulnerable in 2007 (Boertmann 2008).

Polar Bear Table 14.1 The number, sex, and geographic locations of polar bears housed in zoos as of December 31, 2017 (Linke 2017).

Male

Female

Unknown

Total

Institutions

Europe

66

80

5

151

59

North America

31

41

1

73

33

Asia

33

33

0

66

33

Australia

2

2

0

4

1

South America

1

1

0

2

1

Africa

0

0

0

0

0

Total

133

157

6

296

127

Status in Captivity Some of the earliest historical notations of polar bears in human care were polar bears gifted to royalty (i.e. in 894 to King Harald of Norway, and to King Mohammed al Kamul of Egypt in 1234; Engelhard 2017). In 1252, the king of Norway gifted King Henry III of England a polar bear for the Royal Menagerie. In 1693, Fredrick I, King of Prussia used polar bears in mock “Roman arena”-like combat fights with other species. Polar bears were first exhibited in Europe at a zoo in London in 1829 and in North America in 1876 in Philadelphia (Linke 2017). Today, nearly 300 polar bears can be found in zoos on five continents (Table 14.1). In the 1990s many US and European zoos decided to no longer house the species. Although polar bears were still popular with visitors, some zoos and other groups expressed concern about the animals’ welfare and the zoo population was considered to be of low conservation value. In the time between the publication of the first North American regional polar bear studbook in 1995 and the second addition in 2006, the numbers of bears and facilities housing them decreased by almost half (Poirier & Lanthier 1995; Meyerson 2006; Linke 2017). The Association of Zoos and Aquariums (AZA) Species Survival Program (SSP) formed in 2003 with the goal of maintaining a genetically diverse captive population while decreasing the numbers in the AZA population. Population planning for this program included breeding moratoriums and the use of contraception. Animal care and welfare professionals developed new care recommendations, which included input from field biologists and government entities. These recommendations addressed important facility design features, behavioral needs

References Aars, J. (2013). Variation in detection probability of polar bear maternity dens. Polar Biology 36: 1089–1096. Aars, J., Marques, T. A., Lone, K., et al. (2017). The number and distribution of polar bears in the western Barents Sea. Polar Research 36: 1374125. Amstrup, S. C., Marcot, B. G. & Douglas D. C. (2008). A Bayesian network modelling

and enrichment and training programs. In 2006, the European Association of Zoos and Aquaria (EAZA) established the Polar Bear European Endangered Species Programmes (EEP). Similar population trends and welfare concerns were acknowledged by the EAZA Ursid specialists, which led to the development of Ursid Husbandry Guidelines (Meyerson et al. 2017). For more than a decade these two polar bear programs have worked together to enhance the health and welfare of polar bears in human care in North America and Europe. As global warming emerged as the primary threat to polar bears, there has been renewed interest among zoos in housing polar bears as a flagship species for conservation action. With over 180 million people visiting US zoos annually, there is an opportunity to educate the public about the threats facing polar bears and the opportunities for climate change mitigation. More recently, collaborations between zoo and field researchers have demonstrated that bears in human care provide an opportunity for research that informs the status and health of their wild counterparts. Active training programs have already supported the use of polar bears in human care for studies of sensory ecology (olfaction and hearing), energetics, emerging diseases, reproduction, and development of new methods for studying wild bears (Dutton et al. 2009; Bowen et al. 2015; Owen et al. 2015; Ware et al. 2015; Rode et al. 2016; Pagano et al. 2019). The success of these studies led to the creation of the Polar Bear Research Council and the creation of a research masterplan to facilitate research with zoo polar bears that aids in understanding wild populations. This plan was developed by members of the AZA, EAZA, government, non-government, and scientific communities.

approach to forecasting the 21st century worldwide status of polar bears. In: DeWeaver, E. T., Bitz, C. M. & Tremblay, L. B. (Eds.), Arctic sea ice decline: Observations, projections, mechanisms, and implications (pp. 213–268). Geophysical Monograph 180. Washington, DC: American Geophysical Union. Andersen, M., Derocher, A. E., Wiig, Ø. & Aars, J. (2012). Polar bear (Ursus

maritimus) maternity den distribution in Svalbard, Norway. Polar Biology 35: 499–508. Arnould, J. P. Y. & Ramsay, M. A. (1994). Milk production and milk consumption in polar bears during the ice-free period in western Hudson Bay. Canadian Journal of Zoology 72: 1365–1370. Åsbakk, K., Aars, J., Derocher, A. E., et al. (2010). Serosurvey for Trichinella in polar bears (Ursus maritimus) from Svalbard

205

Species Accounts

and the Barents Sea. Veterinary Parasitology 172: 256–263. Atkinson, S. N. & Ramsay, M. A. (1995). The effects of prolonged fasting of the body composition and reproductive success of female polar bears (Ursus maritimus). Functional Ecology 9: 559–567. Atkinson, S. N., Nelson, R. A. & Ramsay, M. A. (1996a). Changes in the body composition of fasting polar bears (Ursus maritimus): the effect of relative fatness on protein conservation. Physiological Zoology 69: 304–316.

Boertmann, D. (2008). Greenland Red List – 2007. Aarhus: Greenland Home Rule. Bowen, L., Miles, A. K., Waters, S., et al. (2015). Gene transcription in polar bears (Ursus maritimus) from disparate populations. Polar Biology 38: 1413–1427.

Atkinson, S. N., Stirling, I. & Ramsay, M. A. (1996b). Growth in early life and relative body size among adult polar bears (Ursus maritimus). Journal of Zoology 239: 225–234.

Bromaghin, J. F., McDonald, T. L., Stirling, I., et al. (2015). Polar bear population dynamics in the southern Beaufort Sea during a period of sea ice decline. Ecological Applications 25: 634–651.

Atwood, T. C., Marcot, B. G., Douglas, D. C., et al. (2015). Evaluating and ranking threats to the long-term persistence of polar bears. US Geological Survey OpenFile Report 2014-1254.

Brown, T. M., Macdonald, R., Muir, D. C. G. & Letcher, R. J. (2018). The distribution and trends of persistent organic pollutants and mercury in marine mammals from Canada’s Eastern Arctic. Science of the Total Environment 618: 500–517.

Atwood, T. C., Peacock, E., McKinney, M. A., et al. (2016). Rapid environmental change drives increased land use by an Arctic marine predator. PLoS ONE 11: e0155932. Atwood, T. C., Duncan, C. G., Patyk, K. A., et al. (2017). Environmental and behavioral changes may influence the exposure of an Arctic apex predator to pathogens and contaminants. Scientific Reports 7: 13193. Bechshøft, T. Ø., Derocher, A. E., Viengkone, M., et al. (2018). On the integration of ecological and physiological variables in polar bear toxicology research: a systematic review. Environmental Review 26: 1-12. Belikov, S. E., Boltunov, A. N., Ovsyanikov, N. G., Mordvintsev, I. & Nikiforov, V. V. (2010). Polar bear management and research in Russia, 2005–2009. In: Obbard, M. E., Thiemann, G. W., Peacock, E. & DeBruyn, T. D. (Eds.), Polar bears: Proceedings of the 15th Working Meeting of the IUCN/SSC Polar Bear Specialist Group, June 29–July 3, 2009, Copenhagen, Denmark. Occasional Paper of the IUCN Species Survival Commission No. 43 (pp. 165–169). Gland, Switzerland and Cambridge, UK: IUCN. Bentzen, T. W., Follmann, E. H., Amstrup, S. C., et al. (2007). Variation in winter diet of southern Beaufort Sea polar bears inferred from stable isotope analysis. Canadian Journal of Zoology 85: 596–608.

206

Bignert, A., Hung, H., Katsoyiannis, A. A., et al. (2016). Arctic Monitoring and Assessment Programme (AMAP) 2016: Temporal Trends in Persistent Organic Pollutants in the Arctic. Oslo, Norway: AMAP.

Cherry, S. G., Derocher, A. E., Thiemann, G. W. & Lunn, N. J. (2013). Migration phenology and seasonal fidelity of an Arctic marine predator in relation to sea ice dynamics. Journal of Animal Ecology 82: 912–921. Clark, D. A., Stirling, I. & Calvert, W. (1997). Distribution, characteristics, and use of earth dens and related excavations by polar bears on the western Hudson Bay Lowlands. Arctic 50: 158–166. Comiso, J. C. (2012). Large decadal decline of the Arctic multiyear ice cover. Journal of Climate 25: 1176–1193. Comiso, J. C., Parkinson, C. L., Gersten, R. & Stock, L. (2008). Accelerated decline in the Arctic sea ice cover. Geophysical Research Letters 35: L01703.

Derocher, A. E. & Stirling, I. (1990a). Distribution of polar bears (Ursus maritimus) during the ice-free period in western Hudson Bay. Canadian Journal of Zoology 68: 1395–1403. Derocher, A. E. & Stirling, I. (1990b). Observations of aggregating behaviour in adult male polar bears (Ursus maritimus). Canadian Journal of Zoology 68: 1390–1394. Derocher, A. E. & Stirling, I. (1994). Agespecific reproductive performance of female polar bears. Journal of Zoology 234: 527–536. Derocher, A. E. & Stirling, I. (1998). Geographic variation in growth of polar bears (Ursus maritimus). Journal of Zoology 245: 65–72. Derocher, A. E. & Wiig, Ø. (2002). Postnatal growth in body length and mass of polar bears (Ursus maritimus) at Svalbard. Journal of Zoology 256: 343–349. Derocher, A. E., Stirling, I. & Andriashek, D. (1992). Pregnancy rates and serum progesterone levels of polar bears in western Hudson Bay. Canadian Journal of Zoology 70: 561–566. Derocher, A. E., Andriashek, D. & Arnould, J. P. Y. (1993). Aspects of milk composition and lactation in polar bears. Canadian Journal of Zoology 71: 561–567. Derocher, A. E., Wiig, Ø. & Andersen, M. (2002). Diet composition of polar bears in Svalbard and the western Barents Sea. Polar Biology 25: 448–452. Derocher, A. E., Andersen, M., Wiig, Ø. & Aars, J. (2010). Sexual dimorphism and the mating ecology of polar bears (Ursus maritimus) at Svalbard. Behavioral Ecology and Sociobiology 64: 939–946. Derocher, A. E., Andersen, M., Wiig, Ø., et al. (2011). Sea ice and polar bear den ecology at Hopen Island, Svalbard. Marine Ecology Progress Series 441: 273–279.

COSEWIC. (2008). COSEWIC Assessment and Update Status Report (WWW document). Available from http:// sararegistry.gc.ca/virtual_sara/files/ cosewic/sr_polar_bear_0808_e.pdf (accessed January 3, 2019).

de Wit, C. A., Muir, D. C. G., Vorkamp, K., et al. (2017). Arctic Monitoring and Assessment Programme (AMAP) 2016: Chemicals of Emerging Arctic Concern. Oslo, Norway: AMAP.

Dau, J. & Barrett, R. (1981). Trichinella. In: Dieterich, R. A. (Ed.), Alaskan wildlife diseases (pp. 151–161). Fairbanks, AK: University of Alaska.

Dietz, R., Sonne, C., Basu, N., et al. (2013). What are the toxicological effects of mercury in Arctic biota? Science of the Total Environment 443: 775–790.

DeMaster, D. P. & Stirling, I. (1981). Ursus maritimus. Mammalian Species 145: 1–7.

Dietz, R., Desforges, J. P., Gustavson, K., et al. (2018). Immunologic, reproductive, and carcinogenic risk assessment from POP exposure in East Greenland polar bears (Ursus maritimus) during

Derocher, A. E. (1999). Latitudinal variation in litter size of polar bears: ecology or methodology? Polar Biology 22: 350–356.

Polar Bear

1983–2013. Environmental International 118: 169–178. Dietz, R., Letcher, R. J., Desforges, J-P., et al. (2019). Current state of knowledge on biological effects from contaminants on arctic wildlife and fish. Science of the Total Environment 696: 133792. Duignan, P. J., Van Bressem, M-F., CortésHinojosa, G. & Kennedy-Stoskopf, S. (2018). Viruses. In: Gulland, F. M. D., Dierauf, L. A. & Whitman, K. L. (Eds.), Handbook of marine mammal diseases, 2nd edition (pp. 331–365). Boca Raton, FL: CRC Press. Duncan, C. G., Tiller, R., Mathis, D., et al. (2014). Brucella placentitis and seroprevalence in northern fur seals (Callorhinus ursinus) of the Pribilof Islands, Alaska. Journal of Veterinary Diagnostic Investigation 26: 507–512. Dupouy-Camet, J., Bourée, P. & Year, H. (2017). Trichinella and polar bears: a limited risk for humans. Journal of Helminthology 91: 440–446. Durner, G. M., Amstrup, S. C. & Fischbach, A. S. (2003). Habitat characteristics of polar bear terrestrial maternal den sites in northern Alaska. Arctic 56: 55–62. Durner, G. M., Douglas, D. C., Nielson, R. M., et al. (2009). Predicting 21st-century polar bear habitat distribution from global climate models. Ecological Monographs 79: 25–58. Durner, G. M., Whiteman, J. P., Harlow, H. J., et al. (2011). Consequences of longdistance swimming and travel over deepwater pack ice for a female polar bear during a year of extreme sea ice retreat. Polar Biology 34: 975–984. Durner, G. M., Laidre, K. L. & York, G. S. (2018). Proceedings of the 18th working meeting of the IUCN/SSC Polar Bear Specialist Group, June 7–11, 2016, Anchorage, AK. Gland, Switzerland and Cambridge, UK: IUCN. Dutton, C. J., Quinnell, M., Lindsay, R., DeLay, J. & Barker, I. K. (2009). Paraparesis in a polar bear (Ursus maritimus) associated with West Nile virus infection. Journal of Zoo and Wildlife Medicine 40: 568–571.

biodiversity/maps-sub-populations-polarbears-protected.html#_fig03 (accessed January 9, 2019). Escajeda, E., Laidre, K. L., Born, E. W., et al. (2018). Identifying shifts in maternity den phenology and habitat characteristics of polar bears (Ursus maritimus) in Baffin Bay and Kane Basin. Polar Biology 41: 87–100. Fagre, A., Patyk, K. A., Nol, P., et al. (2015). A review of infectious agents in polar bears (Ursus maritimus) and their longterm ecological relevance EcoHealth 12: 528–539.

Griffen, B. D. (2018). Modelling the metabolic costs of swimming in polar bears (Ursus maritimus). Polar Biology 41: 491–503. Hamilton, S. G., Castro de la Guardia, L., Derocher, A. E., et al. (2014). Projected polar bear sea ice habitat in the Canadian Arctic Archipelago. PLoS ONE 9: e113746. Ingolfsson, O. &Wiig, Ø. (2009). Late Pleistocene fossil find in Svalbard: the oldest remains of a polar bear (Ursus maritimus Phipps, 1744) ever discovered. Polar Research 28: 455–462.

Ferguson, S. H., Taylor, M. K., Born, E. W., Rosing-Asvid, A. & Messier, F. (1999). Determinants of home range size for polar bears (Ursus maritimus). Ecology Letters 2: 311–318.

Jessen, A. (2018). Management on polar bears in Greenland, 2009–2016. In: Durner, G. M., Laidre, K. L. & York, G. S. (Eds.), Polar bears: Proceedings of the 18th Working Meeting of the IUCN/SSC Polar Bear Specialist Group, Anchorage, Alaska, June 7–11, 2016 (pp. 84–94). Gland, Switzerland and Cambridge, UK: IUCN.

Ferguson, S. H., Taylor, M. K. & Messier, F. (2000). Influence of sea ice dynamics on habitat selection by polar bears. Ecology 81: 761–772.

Kingsley, M. C. S. (1979). Fitting the von Bertalanffy growth equation to polar bear age-weight data. Canadian Journal of Zoology 57: 1020–1025.

Ferguson, S. H., Stirling, I. & McLoughlin, P. (2005). Climate change and ringed seal (Phoca hispida) recruitment in western Hudson Bay. Marine Mammal Science 21: 121–135.

Kingsley, M. C. S., Stirling, I. & Calvert, W. (1985). The distribution and abundance of seals in the Canadian high Arctic, 1980–1982. Canadian Journal of Fisheries and Aquatic Sciences 42: 1189–1210.

Ferguson, S. H., Young, B. G., Yurkowski, D. J., et al. (2017). Demographic, ecological, and physiological responses of ringed seals to an abrupt decline in sea ice availability. Peer Journal 5: e2957.

Kirk, C. M., Amstrup, S., Swor, R., Holcomb, D. & O’Hara, T. M. (2010). Morbillivirus and Toxoplasma exposure and association with hematological parameters for southern Beaufort Sea polar bears: potential response to infectious agents in a sentinel species. EcoHealth 7: 321–331.

Fay, F. H. (1960). Carnivorous walrus and some arctic zoonoses. Arctic 13: 111–122.

Fischbach, A. S., Amstrup, S. C. & Douglas, D. C. (2007). Landward and eastward shift of Alaskan polar bear denning associated with recent sea ice changes. Polar Biology 30: 1395–1405. Freitas, C., Kovacs, K. M., Andersen, M., et al. (2012). Importance of fast ice and glacier fronts for female polar bears and their cubs during spring in Svalbard, Norway. Marine Ecology Progress Series 447: 289–304.

Engelhard, M. (2017). Ice bear. Washington, DC: University of Washington Press.

Galicia, M. P., Thiemann, G. W., Dyck, M. G., Ferguson, S. H. & Higdon, J. W. (2016). Dietary habits of polar bears in Foxe Basin, Canada: possible evidence of a trophic regime shift mediated by a new top predator. Ecology and Evolution 6: 6005–6018.

Environment and Climate Change Canada. (2018). Canadian polar bear subpopulation and status map 2018 (WWW document). Available from www.canada.ca/en/ environment-climate-change/services/

Gormezano, L. J. & Rockwell, R. F. (2013). What to eat now? Shifts in polar bear diet during the ice-free season in western Hudson Bay. Ecology and Evolution 3: 3509–3523.

Knudsen, B. (1978). Time budgets of polar bears (Ursus maritimus) on North Twin Island, James Bay, during summer. Canadian Journal of Zoology 56: 1627–1628. Kolenosky, G. B. & Prevett, J. P. (1983). Productivity and maternity denning of polar bears in Ontario. International Conference on Bear Research and Management 5: 238–245. Kumar, V., Lammers, F., Bidon, T., et al. (2017). The evolutionary history of bears is characterized by gene flow across species. Scientific Reports 7: 46487. Kwok, R. & Rothrock, D. A. (2009). Decline in Arctic sea ice thickness from submarine and ICESat records: 1958–2008. Geophysical Research Letters 36: 2009GL039035. Laidre, K. L., Born, E. W., Heagerty, P., et al. (2015a). Shifts in habitat use by female

207

Species Accounts

polar bears (Ursus maritimus) in East Greenland. Polar Biology 38: 879–893. Laidre, K. L., Stern, H., Kovacs, K.M., et al. (2015b). Arctic marine mammal population status, sea ice habitat loss, and conservation recommendations for the 21st century. Conservation Biology 29: 724–737. Laidre, K. L., Stirling, I., Estes, J. A., Kochnev, A. & Roberts, J. (2018). Historical and potential future importance of large whales as a food for polar bears. Frontiers in Ecology and the Environment 16: 515–524. Larsen, T. S. & Stirling I. (2009). The Agreement on the Conservation of Polar Bears – its History and Future. Norsk Polarinstittut, No. 127. 16 pp. Latour, P. B. (1981). Spatial relationships and behavior of polar bears (Ursus maritimus Phipps) concentrated on land during the ice-free season of Hudson Bay. Canadian Journal of Zoology 56: 1763–1774. Letcher, R. J., Bustnes, J-O., Dietz, R., et al. (2010). Exposure and effects assessment of persistent organic pollutants in Arctic wildlife and fish. Science of the Total Environment 408: 2995–3043. Letcher, R. J., Morris, A. D., Dyck, M., et al. (2018). Legacy and new halogenated persistent organic pollutants in polar bears from a contamination hotspot in the Arctic, Hudson Bay Canada. Science of the Total Environment 610: 121–136. Linke, K. (2017). World Zoo & Aquarium Association International Polar Bear Studbook. Rostock, Germany: Zoo Rostock. Liu, Y., Richardson, E. S., Derocher, A. E., et al. (2018). Hundreds of unrecognized halogenated contaminants discovered in polar bear serum. Angewandte Chemie 57: 16,401–16,406. Loewen, K., Prins, B. & Philibert, H. (1990). Northwest Territories. Rabies in a polar bear. Canadian Veterinary Journal 31: 457. Lone, K., Kovacs, K. M., Lydersen, C., et al. (2018a). Aquatic behaviour of polar bears (Ursus maritimus) in an increasingly ice-free Arctic. Scientific Reports 8: 9677. Lone, K., Merkel, B., Lydersen, C., Kovacs, K. M. & Aars, J. (2018b). Sea ice resource selection models for polar bears in the Barents Sea subpopulation. Ecography 41: 567–578.

208

Lunn, N. J., Stirling, I., Andriashek, D. & Richardson, E. (2004). Selection of maternity dens by female polar bears in western Hudson Bay, Canada and the effects of human disturbance. Polar Biology 27: 350–356. Lunn, N. J., Servanty S., Regehr, E.V., et al. (2016). Demography of an apex predator at the edge of its range: impacts of changing sea ice on polar bears in Hudson Bay. Ecological Applications 26: 1302–1320. Malenfant, R. M., Coltman, D. W., Richardson, E. S., et al. (2015). Evidence of adoption, monozygotic twinning, and low inbreeding rates in a large genetic pedigree of polar bears. Polar Biology 39: 1455–1465. Malenfant, R. M., Davis, C. S., Cullinghman, C. I. & Coltman, D. W. (2016). Circumpolar genetic structure and recent gene flow of polar bears: a reanalysis. PLoS ONE 11: e0148967. Manly, B. F. J., McDonald, L. L., Thomas, D. L., McDonald, T. L. & Erickson, W. P. (2002). Resource selection by animals. Dordrecht: Kluwer Academic Publishers. Mauritzen, M., Derocher, A. E., Wiig, Ø., et al. (2002). Using satellite telemetry to define spatial population structure in polar bears in the Norwegian and western Russian Arctic. Journal of Applied Ecology 39: 79–90. Mauritzen, M., Belikov, S. E., Boltunov, A. N., et al. (2003). Functional responses in polar bear habitat selection. Oikos 100: 112–124.

exposures of East Greenland polar bears. Global Change Biology 19: 2360–2372. McKinney, M. A., Pedro, S., Dietz, R., et al. (2015). Ecological impacts of global climate change on persistent organic pollutant and mercury pathways and exposures in arctic marine ecosystems: a review of initial findings. Current Zoology 61: 617–628. Messier, F., Taylor, M. K. & Ramsay, M. A. (1992). Seasonal activity patterns of female polar bears (Ursus maritimus) in the Canadian Arctic as revealed by satellite telemetry. Journal of Zoology 226: 219–229. Messier, F., Taylor, M. K. & Ramsay, M. A. (1994). Denning ecology of polar bears in the Canadian Arctic archipelago. Journal of Mammalogy 75: 420–430. Messier, V., Levesque, B., Proulx, J. F., et al. (2009). Seroprevalence of Toxoplasma gondii among Nunavik Inuit (Canada). Zoonoses Public Health 56: 188–197. Meyerson, R. (2006). Association of Zoo and Aquariums Polar Bear Studbook. Toledo, OH: Toledo Zoo. Meyerson, R., Moore, D., Long, S. & CheCastaldo, J. (2017). Welfare of captive polar bears and their value in in situ conservation efforts. In: Butterworth, A. (Ed.), Marine mammal welfare – Humaninduced change in the marine environment and its impacts on marine mammal welfare (pp. 489–502). Bristol, UK: Springer Publishing.

McCall, A. G., Pilfold, N. W., Derocher, A. E. & Lunn, N. J. (2016). Seasonal habitat selection by adult female polar bears in western Hudson Bay. Population Ecology 58: 407–419.

Miller, M., Shapiro, K., Murray, M. J., Haulena, M. & Raverty, S. (2018). Protozoan parasites of marine mammals. In: Gulland, F. M. D., Dierauf, L. A. & Whitman, K. L. (Eds.), Handbook of marine mammal diseases, 3rd edition (pp. 425–469). Boca Raton, FL: CRC Press.

McDonald, J. C., Gyorkos, T. W., Alberton, B., et al. (1990) An outbreak of toxoplasmosis in pregnant women in northern Quebec. Journal of Infectious Diseases 161: 769–774.

Miller, S., Wilder, J. & Wilson, R. R. (2015). Polar bear–grizzly bear interactions during the autumn open-water period in Alaska. Journal of Mammalogy 96: 1317–1325.

McKinney, M. A., Stirling, I., Lunn, N. J., Peacock, E. & Letcher, R. J. (2010). The role of diet on long-term concentration and pattern trends of brominated and chlorinated contaminants in western Hudson Bay polar bears, 1991–2007. Science of the Total Environment 408: 6210–6222.

Miller, W., Schuster, S. C., Welch, A. J., et al. (2012). Polar and brown bear genomes reveal ancient admixture and demographic footprints of past climate change. Proceedings of the National Academy of Sciences 109: 2382–2390.

McKinney, M. A., Iverson, S., Fisk, A., et al. (2013). Global change effects on the longterm feeding ecology and contaminant

Molnár, P. K., Derocher, A. E., Lewis, M. A. & Taylor, M. K. (2008). Modelling the mating system of polar bears: a mechanistic approach to the Allee effect.

Polar Bear Proceedings of the Royal Society B – Biological Sciences 275: 217–226. Monnett, C. & Gleason, J. S. (2006). Observations of mortality associated with extended open-water swimming by polar bears in the Alaskan Beaufort Sea. Polar Biology 29: 681–687. Nymo, I., Tryland, M., Kirstine-Frie, A., et al. (2013). Age-dependent prevalence of anti-Brucella antibodies in hooded seals Cystophora cristata. Inter-Research 106: 187–196.

polar bears in the eastern Beaufort Sea. PLoS ONE 7: e41429. Pilfold, N. W., Derocher, A. E. & Richardson, E. (2014). Influence of intraspecific competition on the distribution of a wide-ranging, nonterritorial carnivore. Global Ecology and Biogeography 23: 425–435. Poirier, P. & Lanthier, C. (1995). North American Regional Polar Bear Studbook. Quebec, Canada: Jardin zoologique de Granby.

Obbard, M. E., Cattet, M. R. L., Howe, E. J., et al. (2016). Trends in body condition in polar bears (Ursus maritimus) from the Southern Hudson Bay subpopulation in relation to changes in sea ice. Arctic Science 2: 15–32.

Pond, C. M., Mattacks, C. A., Colby, R. H. & Ramsay, M. A. (1992). The anatomy, chemical composition, and metabolism of adipose tissue in wild polar bears (Ursus maritimus). Canadian Journal of Zoology 70: 326–341.

O’Hara, T. M., Holcomb, D., Elzer, P., et al. (2010). Brucella species survey in polar bears (Ursus maritimus) of northern Alaska. Journal of Wildlife Diseases 46: 687–694.

Pongracz, J. D., Paetkau, D., Branigan, M. & Richardson, E. (2017). Recent hybridization between a polar bear and grizzly bears in the Canadian Arctic. Arctic 70: 151–160.

Olson, J. W., Rode, K. D., Eggett, D. L., et al. (2017). Collar temperature sensor data reveal long-term patterns in southern Beaufort Sea polar bear den distribution on pack ice and land. Marine Ecology Progress Series 564: 211–224.

Prop, J., Aars, J., Bardsen, B.-J., et al. (2015). Climate change and the increasing impact of polar bears on bird populations. Frontiers in Ecology and Evolution 3: 33.

Owen, M. A., Swaisgood, R. R., Slocomb, C., et al. (2015). An experimental investigation of chemical communication in the polar bear. Journal of Zoology 295: 36–43. Pagano, A., Cutting, A., Nicassio-Hiskey, N., Hash, A. & Williams, T. (2019). Energetic costs of aquatic locomotion in a sub-adult polar bear. Marine Mammal Science 35, 649–659. Parks, E. K., Derocher, A. E. & Lunn, N. J. (2006). Seasonal and annual movement patterns of polar bears on the sea ice of Hudson Bay. Canadian Journal of Zoology 84: 1281–1294. Peacock, E., Taylor, M. K., Laake, J. L. & Stirling, I. (2013). Population ecology of polar bears in Davis Strait, Canada and Greenland. Journal of Wildlife Management 77: 463–476. Philippa, J. D., Leighton, F. A., Daoust, P. Y., et al. (2004). Antibodies to selected pathogens in free-ranging terrestrial carnivores and marine mammals in Canada. Veterinary Record 155: 135–140. Pilfold, N. W., Derocher, A. E., Stirling, I., Richardson, E. & Andriashek, D. (2012). Age and sex composition of seals killed by

Rah, H., Chomel, B. B., Kasten, R. W., et al. (2005) Serosurvey of selected zoonotic agents in polar bears (Ursus maritimus). Veterinary Record 156: 7–13. Ramsay, M. A. & Dunbrack, R. L. (1986). Physiological constraints on life-history phenomena: the example of small bear cubs at birth. American Naturalist 127: 735–743. Ramsay, M. A. & Stirling, I. (1986). On the mating system of polar bears. Canadian Journal of Zoology 64: 2142–2151. Ramsay, M. A. & Stirling, I. (1988). Reproductive biology and ecology of female polar bears (Ursus maritimus). Journal of Zoology 214: 601–634. Ramsay, M. A. & Stirling, I. (1990). Fidelity of female polar bears to winter den sites. Journal of Mammalogy 71: 233–236. Regehr, E. V., Lunn, N. J., Amstrup, S. C. & Stirling, I. (2007). Effects of earlier sea ice breakup on survival and population size of polar bears in western Hudson Bay. Journal of Wildlife Management 71: 2673–2683. Regehr, E. V., Hunter, C. M., Caswell, H., Amstrup, S. C. & Stirling, I. (2010). Survival and breeding of polar bears in the southern Beaufort Sea in relation to sea ice. Journal of Animal Ecology 79: 117–127.

Regehr, E. V, Laidre, K. L., Akçakaya, H. R., et al. (2016). Conservation status of polar bears (Ursus maritimus) in relation to projected sea-ice declines. Biology Letters 12: 20160556. Regehr, E. V., Hostetter, N. J., Wilson, R. R., et al. (2018). Integrated population modeling provides the first empirical estimates of vital rates and abundance for polar bears in the Chukchi Sea. Scientific Reports 8: 16780. Richardson, E., Stirling, I. & Hik, D. S. (2005). Polar bear (Ursus maritimus) maternity denning habitat in western Hudson Bay: a bottom-up approach to resource selection functions. Canadian Journal of Zoology 83: 860–870. Rigét, F. F., Bignert, A., Braune, B., et al. (2019). Temporal trends of persistent organic pollutants in Arctic marine and freshwater biota. Science of the Total Environment 649: 99–110. Robbins, C. T., Ben-David, M., Fortin, J. K. & Nelson, O. L. (2012). Maternal condition determines birth date and growth of newborn bear cubs. Journal of Mammalogy 93: 540–546. Rode, K. D., Amstrup, S. C. & Regehr, E. V. (2010). Reduced body size and cub recruitment in polar bears associated with sea ice decline. Ecological Applications 20: 768–782. Rode, K. D., Regehr, E. V., Douglas, D. C., et al. (2014).Variation in the response of an Arctic top predator experiencing habitat loss: feeding and reproductive ecology of two polar bear populations. Global Change Biology 20: 76–88. Rode, K. D., Wilson, R. R., Regehr, E. V., et al. (2015a). Increased land use by Chukchi Sea polar bears in relation to changing sea ice conditions. PLoS ONE 10: e0142213. Rode, K. D., Robbins, C. T., Nelson, L., et al. (2015b). Can polar bears use terrestrial foods to offset lost ice-based hunting opportunities? Frontiers in Ecology and the Environment 13: 138–145. Rode, K. D., Stricker, C. A., Erlenbach, J., et al. (2016). Isotopic incorporation and the effects of fasting and dietary lipid content on isotopic discrimination in large, carnivorous mammals. Physiological and Biochemical Zoology 89: 182–197. Rosing-Asvid, A., Born, E. W. & Kingsley, M. C. S. (2002). Age at sexual maturity of males and timing of the mating season of

209

Species Accounts

polar bears (Ursus maritimus) in Greenland. Polar Biology 25: 878–883. Routti, H., Atwood, T., Bechshoft, T. Ø., et al. (2019). State of knowledge on current exposure, fate and potential health effects of contaminants in polar bears from the circumpolar Arctic. Science of the Total Environment 664, 1063–1083. Russell, R. H. (1975). The food habits of polar bears of James Bay and southwest Hudson Bay in summer and autumn. Arctic 28: 117–129. Schliebe, S., Rode, K. D., Gleason, J. S., et al. (2008). Effects of sea ice extent and food availability on spatial and temporal distribution of polar bears during the fall openwater period in the Beaufort Sea. Polar Biology 31: 999–1010. Sidor, I., Goldstein, T. & Whatmore, A. M. (2008). Transplacental infection with Brucella in California sea lion (Zalophus californianus) fetus. In Brucellosis International Research Conference, Chicago, USA. Smith, T. G. & Stirling, I. (1975). The breeding habitat of the ringed seal (Phoca hispida). The birth lair and associated structures. Canadian Journal of Zoology 53: 1297–1305. Smith, T. G. & Stirling, I. (1978). Variation in the density of ringed seal (Phoca hispida) birth lairs in the Amundsen Gulf, Northwest Territories. Canadian Journal of Zoology 56: 1066–1070. Smith, T. S., Partridge, S. T., Amstrup, S. C. & Schliebe, S. (2007). Post-den emergence behavior of polar bears (Ursus maritimus) in northern Alaska. Arctic 60: 187–194. Sonne, C. (2010). Health effects from longrange transported contaminants in Arctic top predators: an integrated review based on studies of polar bears and relevant model species. Environment International 36: 461–491. Sonne, C., Letcher, R. J., Bechshøft, T. Ø., et al. (2012). Two decades of biomonitoring polar bear health in Greenland: a review. Acta Veterinaria Scandinavica 54: S15. Sonne, C., Bechshøft, T. Ø. & Rigét, F. F. (2013). Size and density of East Greenland polar bear (Ursus maritimus) skulls: valuable bioindicators of environmental changes? Ecological Indicators 34: 290–295.

210

Stern, H. L. & Laidre, K. L. (2015). Sea-ice indicators of polar bear habitat. The Cryosphere 10: 2027–2041. Stirling, I. (1974). Midsummer observations on the behavior of wild polar bears (Ursus maritimus). Canadian Journal of Zoology 52: 1191–1198. Stirling, I. (1997). The importance of polynyas, ice edges, and leads to marine mammals and birds. Journal of Marine Systems 10: 9–21. Stirling, I. & Archibald, W. R. (1977). Aspects of predation on seals by polar bears. Journal of Fisheries Research Board Canada 34: 1126–1129. Stirling, I. & Latour, P. B. (1978). Comparative hunting abilities of polar bear cubs of different ages. Canadian Journal of Zoology 56: 1768–1772. Stirling, I. & van Meurs, R. (2015). Longest recorded underwater dive by a polar bear. Polar Biology 38: 1301–1304. Stirling, I., Kingslay, M. C. S. & Calvert, W. (1982). The distribution and abundance of seals in the eastern Beaufort Sea, 1974–79. Canadian Wildlife Service Occasional Report 46. Stirling, I., Andriashek, D. & Calvert, W. (1993). Habitat preferences of polar bears in the western Canadian Arctic in late winter and spring. Polar Record 29: 13–24. Stirling, I., Lunn, N. J. & Iacozza, J. (1999). Long-term trends in the population ecology of polar bears in western Hudson Bay in relation to climate change. Arctic 52: 294–306. Stirling, I., Lunn, N. J., Iacozza, J., Elliott, C. & Obbard, M. (2004). Polar bear distribution and abundance on the southwestern Hudson Bay coast during open water season, in relation to population trends and annual ice patterns. Arctic 57: 15–26. Stirling, I., McDonald, T. L., Richardson, E. S., Regehr, E. V. & Amstrup, S. C. (2011). Polar bear population status in the northern Beaufort Sea, Canada, 1971–2006. Ecological Applications 21: 859–876. Stirling, I., Spencer, C. & Andriashek, D. (2016). Behavior and activity budgets of wild breeding polar bears (Ursus maritimus). Marine Mammal Science 32: 13–37. Stishov, M. S. (1997). Results of aerial counts of the polar bear dens on the Arctic coast of the extreme Northeast Asia.

Proceedings of the working meeting of the IUCN polar bear specialist group. International Union for the Conservation of Nature Species Survival Commission Occasional Paper 7: 90–92. SWG (Scientific Working Group to the Canada-Greenland Joint Commission on Polar Bear). (2016). Re-assessment of the Baffin Bay and Kane Basin polar bear subpopulations: final report to the Canada–Greenland Joint Commission on polar bear. July 31, 2016. Tartu, S., Aars, J., Andersen, M., et al. (2018). Choose your poison – space-use strategy influences pollutant exposure in Barents Sea polar bears. Environmental Science and Technology 52: 3211–3221. Taylor, M., Elkin, B., Maier, N. & Bradley, M. (1991). Observation of a polar bear with rabies. Journal of Wildlife Diseases 27: 337–339. Thiemann, G., Iverson, S. J. & Stirling, I. (2007). Variability in the blubber fatty acid composition of ringed seals (Phoca hispida) across the Canadian Arctic. Marine Mammal Science 23: 241–261. Thiemann, G. W., Iverson, S. J. & Stirling, I. (2008). Polar bear diets and arctic marine food webs: insights from fatty acid analysis. Ecological Monographs 78: 591–613. Thorshaug, K. & Rosted, A. F. (1956). Researches into the prevalence of trichinosis in animals in Arctic and Antarctic waters. Nordic Veterinary Medicine 8: 115–129. Tryland, M. (2000). Zoonoses of arctic mammals. Infectious Disease Review 2: 55–64. Tryland, M., Derocher, A. E., Wiig, Ø. & Godfroid, J. (2001). Brucella sp. antibodies in polar bears from Svalbard and the Barents Sea. Journal of Wildlife Diseases 37: 523–531. Tryland, M., Neuvonen, E., Huovilainen, A., et al. (2005). Serologic survey for selected virus infections in polar bears at Svalbard. Journal of Wildlife Diseases 41: 310–316. Tryland, M., Nesbakken, T., Robertson, L., Grahek-Ogden, D. & Lunestad, B. T. (2014). Human pathogens in marine mammal meat – a Northern perspective. Zoonoses and Public Health 60: 1–18. Tryland, M., Larsen, A. K. & Nymo, I. H. (2018). Bacterial infections and diseases. In Gulland, F. M. D., Dierauf, L. A. & Whitman, K. L. (Eds.), Handbook of marine mammal diseases, 2nd edition

Polar Bear

(pp. 367–388). Boca Raton, FL: CRC Press. US Fish and Wildlife Service. (2008). Final Rule Listing the Polar Bear as a Threatened Species Under the Endangered Species Act (May 15, 2008) www.fws.gov/alaska/ fisheries/mmm/polarbear/esa.htm. Accessed 3 January 2019. Vongraven, D., Ekker, M. & Wiig, Ø. (2010). Management of polar bears in Norway, 2005–2009. In: Obbard, M. E., Thiemann, G. W., Peacock, E. & DeBruyn, T. D. (Eds.), Polar bears: Proceedings of the 15th Working Meeting of the IUCN/SSC Polar Bear Specialist Group, June 29–July 3, 2009, Copenhagen, Denmark. Occasional Paper of the IUCN Species Survival Commission No. 43 (pp. 149–155). Gland, Switzerland and Cambridge, UK: IUCN. Ware, J. V., Rode, K. D., Pagano, A. M., et al. (2015). Validation of mercury tip-switch and accelerometer activity sensors for

identifying resting and active behavior in bears. Ursus 26: 8–18. Ware, J. V., Rode, K. D., Bromaghin, J. F., et al. (2017). Habitat degradation affects the summer activity of polar bears. Oecologia 184: 87–99. Whiteman, J. P., Harlow, H. J., Durner, G. M., et al. (2015). Summer declines in activity and body temperature offer polar bears limited energy savings. Science 349: 295–298. Wiig, Ø., Gjertz, I., Hansson, R. & Thomassen, J. (1992). Breeding behaviour of polar bears in Hornsund, Svalbard. Polar Record 28: 157–159. Wiig, Ø., Born, E. W. & Pedersen, L. T. (2003). Movements of female polar bears (Ursus maritimus) in the East Greenland pack ice. Polar Biology 26: 509–516. Wiig, Ø., Amstrup, S., Atwood, T., et al. (2015). Ursus maritimus. The IUCN Red List of Threatened Species e.T22823A1.

Wilson, R. R., Horne, J. S., Rode, K. D., Regehr, E. V. & Durner, G. M. (2014). Identifying polar bear resource selection patterns to inform offshore development in a dynamic and changing Arctic. Ecosphere 5: 136. Wimsatt, W. A. (1974). Delayed implantation in the Ursidae, with particular reference to the black bear (Ursus americanus, Pallas). In: Enders, A. C. (Ed.), Delayed implantation (pp. 49–86). Chicago, IL: University of Chicago Press. Zeyl, E., Aars, J., Ehrich, D., Bachmann, L. & Wiig, Ø. (2009). The mating system of polar bears: a genetic approach. Canadian Journal of Zoology 87: 1195–1209. Zeyl, E., Ehrich, D., Aars, J., Bachmann, L. & Wiig, Ø. (2010). Denning-area fidelity and mitochondrial DNA diversity of female polar bears (Ursus maritimus) in the Barents Sea. Canadian Journal of Zoology 88: 1139–1148.

211

Part III Chapter

15

Human–Bear Coexistence

Human–Bear Conflicts at the Beginning of the Twenty-First Century: Patterns, Determinants, and Mitigation Measures Miha Krofel, Marcus Elfstro¨m, Hu¨seyin Ambarlı, Giulia Bombieri, Enrique Gonza´lez-Bernardo, Klemen Jerina, Andre´s Laguna, Vincenzo Penteriani, James P. Phillips, Nuria Selva, Seth M. Wilson, Alejandra Zarzo-Arias, Claudio Groff, Djuro Huber, Alexandros A. Karamanlidis, Yorgos Mertzanis, Eloy Revilla, and Carlos Bautista

Introduction Conflicts between humans and bears have occurred since prehistory (Zedrosser et al. 2011). Through time, the catalogue of human–bear conflicts (HBC) has been changing depending on the values and needs of human societies and their interactions with bears. Even today, conflict situations vary among the eight species of bears and geographically across these species’ ranges (Can et al. 2014). It therefore comes as no surprise that there is no single global definition of HBC accepted among world practitioners and researchers. For the purposes of this book, HBCs were defined as any interaction between humans and bears that result in negative impacts on human livelihood in terms of health, economy, and sociocultural aspects. We do, however, recognize that HBCs also importantly affect bear populations and represent a major threat to bear conservation worldwide. Conflicts may reduce human tolerance to bears, increase poaching, undermine conservation efforts, and increase demands for legal management removals. Successful conflict mitigation is therefore not only important for people’s well-being, but also represents a crucial part of the modern conservation management of bears on a global scale. Most bear species are opportunistic omnivores, they are all large-bodied and potentially dangerous to humans. They live in diverse biomes and come in contact with various human societies. This results in a broad range of interactions between bears and humans that may be considered as conflicts, including: (1) predation of domestic or semi-wild animals, including bees, hunting dogs, and pet animals; (2) damage due to foraging on cultivated berries, fruits, agricultural products, and tree bark in forest plantations; (3) economic loss due to destruction of beehives, fences, silos, houses, and other human property; (4) bear attacks on humans causing mild or fatal trauma; (5) bluff charges, bear intrusions into residential areas and direct proximity of humans or other forms of bear behavior that provoke strong unease, fear, and/or defensive reaction by humans; and (6) vehicle collisions with bears and traffic accidents indirectly provoked by bears during “bear jams” (Can et al. 2014; Bautista et al. 2017).

In this chapter we aim to outline the principal types of HBC and geographical differences in the occurrence of conflicts, show direct and indirect causes of conflicts, illuminate the human dimensions of HBC, and review methods that are used to mitigate HBC with several successful examples that resulted in improved coexistence between people and bears.

Overview of Types of Human–Bear Conflicts Bear conflict literature, reviewed from 1970 to 2018, showed that: (1) HBCs have never been reported for giant pandas; (2) brown and black bears are the species involved in most types of conflicts; whereas (3) Andean, sun, sloth, and polar bears are involved in few, specific types of HBC (Figure 15.1). Some bear species can also represent a threat to human safety, but a specific chapter (Chapter 17) on bear attacks on humans is included in this volume and therefore they are not discussed here. In this section, we present an overview of the different types of conflicts grouped by the bear species for which conflicts have been recorded.

Andean Bear Although Andean bear diet is mostly based on plants, this species will attack domestic animals if available in their habitat (Zukowski & Ormsby 2016; Borbón-García et al. 2017). Cattle and sheep are the most common livestock animals in the Andean bear distribution range and are also most often attacked, although also attacks on goats and horses have been reported (Goldstein 1991; Jorgenson & Sandoval 2005; ZapataRíos & Branch 2018). In recent years, the number of damages to livestock has increased as the best grazing lands coincide with the best bear habitats and an increasing number of locals are changing from growing crops to breeding cattle, especially for dairy (Jampel 2016; Zukowski & Ormsby 2016). Occasionally Andean bears raid crops (especially corn), mainly when fields are close to forest cover and away from human settlements (Jorgenson & Sandoval 2005; Espinosa & Jacobson 2012; Jampel 2016).

213

Human–Bear Coexistence

Research on HBCs involving sloth bears is scarce and focuses mainly on bear attacks on humans (Can et al. 2014; Debata et al. 2017), which is the most frequently reported type of conflict (further addressed in Chapter 17). Although these bears rarely approach human areas, they can cause damage to agricultural fields when seeking crops (such as corn, sweet potatoes, potatoes, onions, or groundnuts; Rajpurohit & Krausman 2000; Bargali et al. 2005; Garcia et al. 2016).

various horticultures (Jonker et al. 1998; Baruch-Mordo et al. 2008; Treves et al. 2010; Shivik et al. 2011; Ditmer et al. 2015). Despite being less adapted to predation on large mammals compared to other large carnivores, American black bears also attack and kill livestock (Garshelis 2009). Most common victims are domestic sheep, goats, pigs, cattle, horses, and other smaller mammals and birds (Baruch-Mordo et al. 2008; Linnell et al. 2013). Apiaries are another strong attractant for black bears, which may cause considerable damage to beehives (Clark et al. 2005; McKinley et al. 2014) to obtain a highenergy food source (Beckmann & Berger 2003). American black bears can also cause considerable losses to forest landowners due to feeding or tree marking (rubbing). For example, bears strip bark from coniferous trees during spring to feed on newly forming vascular tissues for their relatively high content of free-floating sugars. This may result in significant damage (Kimball et al. 1998) and is mainly concentrated on younger stands (Stewart et al. 1999; Ziegltrum & Nolte 2001). They regularly seek out garbage, pet food, bird feeders, and other anthropogenic food and damage human property in the process, for example by breaking into houses, barns, vehicles, and other facilities (Miller & Tutterrow 1999; Breck et al. 2008, 2009; Treves et al. 2010). Other types of damage include bear–vehicle collisions and attacks on pets (Baruch-Mordo et al. 2008; Lowery et al. 2012; Wynn-Grant et al. 2018).

Asiatic Black Bear

Brown Bear

Conflicts between humans and Asiatic black bears have been reported across the whole species’ range in Asia, including China (Liu et al. 2011; Huang et al. 2018), Japan (Saito et al. 2008), Pakistan (Ahmad et al. 2016; Ali et al. 2018), Iran (Ghadirian et al. 2017), India (Chauhan 2003), and Bhutan (Sangay & Vernes 2008; Jamtsho & Wangchuk 2016). The most frequently reported type of conflict is crop-raiding, especially of corn, potatoes, and various fruit trees (Figure 15.1). This kind of damage is concentrated during the summer (Charoo et al. 2011; Scotson et al. 2014), whereas attacks on livestock have been mostly documented during the autumn (Li et al. 2013; Mir et al. 2015; Jamtsho & Wangchuk 2016). Rarely, Asiatic black bears also raid beehives and fish farms (Liu et al. 2011; Jamtsho & Wangchuk 2016). Bark stripping has been reported in Japan, where bears are considered to be one of the greatest sources of conflict with forestry due to damage to tree trunks and consequent reduction of their value (Kobashikawa & Koike 2016).

Conflicts between people and brown bears are very diverse and are mainly related to the bears’ opportunistic foraging and consumption of food. Across its range, the brown bear is the species most often reported to cause damage to livestock (Figure 15.1), which ranges from rabbits and chickens to cattle. Usually livestock damage is highest when brown bears occur in areas with sheep farming (Kaczensky 1999; Gunther et al. 2004). Besides livestock, brown bears have been reported killing pets, captive game animals, and fish (Molinari et al. 2016). When searching for insect larvae, brown bears also frequently destroy beehives (Figure 15.2), which represent the major HBC in some parts of the species’ range (e.g. Groff et al. 2010; Karamanlidis et al. 2011; Naves et al. 2018). Damage in agriculture includes crop fields (especially corn), gardens, orchards, grass silage, and vineyards (Krofel & Jerina 2012). In some parts of the range, brown bears cause damage to forestry, which seems to be concentrated on mature conifer trees (Zyśk-Gorczyńska et al. 2016). Due to their large size, brown bears can also cause considerable damage to buildings, vehicles, wildlife feeders, and other human property while searching for human foods or during vehicle collisions (Krofel & Jerina 2012). Searching for anthropogenic food sources can also bring bears close to human settlements, which increases people’s concern for human safety and often results in lethal removal of the intruding bear (Gunther et al. 2004).

Sun Bear Although little information is available on conflicts with sun bears, this species has been reported causing damage to crops and, more rarely, chickens and goats (Fredriksson 2005; Scotson et al. 2014; Wong et al. 2015). In Indonesian Borneo, sun bears have been reported damaging palm oil plantations and coconut trees by eating the fruits and new growth shoots (Fredriksson 2005). Sun bears are also known to feed on a variety of other fruits and horticultural products in mixed orchards such as corn, snakefruit, jackfruit, sugarcane, banana, pineapple, pumpkin, papaya, and oil palm fruits (Normua et al. 2004; Fredriksson 2005; Sethy & Chauhan 2013; Guharajan et al. 2017), as well as causing damage to farmhouses in search of household products such as rice, sugar, and palm oil (Wong et al. 2015; Figure 15.1).

Sloth Bear

American Black Bear American black bears frequently raid crops, which represent an easy and calorically rich foraging opportunity (Mazur & Seher 2008; Merkle et al. 2013; Ditmer et al. 2015). Most frequently reported crop damage refers to cereals (Shivik et al. 2011; Ditmer et al. 2015), silage, corn, orchards, and

214

Human–Bear Conflicts

Brown bear American black bear Sun bear Sloth bear Polar bear Andean bear Asiatic black bear

Number or references

30

20

10

0

e

ag

ns io lis ad col Ro lls/ ts ki pe on

s ck ta At

e

ag

m

da

am

td

s re Fo

s ive e eh ag Be m da ty an er m op hu Pr g in ch ts oa en pr m Ap ttle e se ag

m

da

ck to

s ve

p ro

C

Li

Types of human–bear conflicts Figure 15.1 Types of human–bear conflicts reported in the scientific literature between 1970 and 2018. For each type of human–bear conflict, the number of related publications is shown. (Bear image credits: the image was downloaded from www.123rf.com, Image ID 107801431, copyright: vastard.) Figure 15.2 Human–bear conflicts are very diverse, but mostly associated with bear foraging behavior, such as raiding beehives to obtain protein-rich bee larvae. (Photo by Miha Krofel.)

215

Human–Bear Coexistence

Polar Bear Damage caused by polar bears is mainly related to human food stores, vehicles, attacks on domestic dogs, and entry into human settlements (Stenhouse et al. 1988; Dyck 2006). This damage derives from the bears’ search for food, and is typically concentrated in August–November, when bears are on land due to melting sea ice (Stenhouse et al. 1988; Dyck 2006; Towns et al. 2009) and do not have access to the seals, their main prey (DeMaster & Stirling 1981).

Main Drivers and Factors That Influence Conflict Occurrence Intrinsic Drivers Large carnivores often avoid human activity and settlements. However, when large carnivores do occur close to people and settlements, it is often interpreted as a behavioral response, such as tolerance to human presence, and/or an association between humans and attractive foods. The waning of a response to a stimulus (e.g. human presence) can explain the increased tolerance toward humans in animals with increasing non-threatening encounters with people, and is hereafter called human habituation. Food resources are unconditioned stimuli, i.e. not related with other stimuli, which usually results in the response of foraging. However, feeding can become conditioned upon an unrelated stimulus, such as human presence or settlements, after repeated association between human activity and food resources (Immelmann & Beer 1989); this process is hereafter called food conditioning. These learning processes can explain why some animals (habituated ones) are not wary of humans or often occur near settlements, while other conspecifics (nonhabituated) still avoid them. However, human habituation and food conditioning do not predict that exposures to stimuli will vary among conspecifics (e.g. sex/age or reproductive categories). Bears can also transmit human tolerance by observational learning from mother to offspring, i.e. by social or cultural transmission (Morehouse et al. 2016). Similarly, young bears may become food-conditioned through their mother’s behavior. However, regardless of the influence of cultural transmission, the development of positive associations between bears and human-derived foods requires some earlier experience with, or cues from, people, human activity, or settlements. In contrast to habituation and conditioning, if animals occur near people because they lack cumulative experience of them (Bejder et al. 2009), i.e. are naïve, this would predict a higher frequency of younger individuals near human activity and settlements. Thus, subadult bears may approach people or settlements due to their naïvety. Human habituation and food conditioning are common responses by bears due to frequent exposure to people and human-derived foods (McCullough 1982; Herrero et al. 2005). However, human activity or settlements may also provide a

216

refuge (sometimes termed human shield) for young bears and females with young against dominant conspecifics (Wielgus & Bunnell 1994; Mueller et al. 2004; Steyaert et al. 2013). Actually, predation and aggression towards conspecific young (especially cubs-of-the-year and yearlings) by older individuals, especially males, constitute a common pattern in brown bears (Swenson et al. 1997, 2001; McLellan 2005). A common pattern across Europe, North America, and Japan is that brown bears, polar bears, and American and Asiatic black bears occurring near human activity and settlements are typically younger, and females accompanied by their offspring are more often near settlements or humans than adult males or lone adult females (Elfström et al. 2014c). The pattern of dominant bears displacing subordinate conspecifics from areas with higher habitat quality in terms of food and cover is referred to as a despotic distribution (Elfström et al. 2014c). In some areas, these displaced bears occurring near settlements often become human-habituated or foodconditioned (Mattson et al. 1992; Schwartz et al. 2006). In other areas, in particular among Scandinavian brown bears, food-conditioning seems to be uncommon (Elfström et al. 2014a, 2014b). Thus, the despotic distribution explains why certain bears more often occur near settlements, which may or may not result in human habituation or food conditioning. The availability of food resources and their distribution in relation to human settlements can differ considerably across the bears’ range and this could explain why sometimes bears near settlements become food-conditioned, but not always.

Extrinsic Drivers Based on a literature review we identified six main extrinsic drivers of HBC: availability of anthropogenic food, natural food failures, impacts of human activities on the landscape, climatic and meteorological events, differences in conflict management, and reduced human tolerance. Below, we give an overview of the mechanisms underlying these drivers, except for the reduced tolerance that is explained further in this chapter. At a local scale, the availability of anthropogenic food (Figure 15.3) is a source of conflict and an ecological trap for bears in many parts of the world, for example: brown bears roaming in agriculture fields or garbage dumps in Japan, Europe, the Middle East, and North America (Narita et al. 2011; Northrup et al. 2012; Cozzi et al. 2016; Penteriani et al. 2018a); American black bears raiding mango plantations in Northern Mexico (Lira Torres et al. 2015); sun and Asiatic black bears raiding corn plantations in south-east Asia (Scotson et al. 2014); and Andean bears predating free-ranging cattle in pastures of the Ecuadorian Andes (Jampel 2016). Additionally, artificial feeding of bears, both intentionally and unintentionally, may increase the risk of HBC, which is why intentional bear feeding is generally discouraged or prohibited across North America (Garshelis et al. 2017). Artificial feeding can also shorten the hibernation period and thus

Human–Bear Conflicts

Figure 15.3 Anthropogenic food, such as garbage, is one of the most important extrinsic drivers of human–bear conflicts. (Photo by Andrej Sila, Slovenia Forest Service.)

prolong the period when human–bear interactions occur (Krofel et al. 2017). The frequency of bears using or searching for anthropogenic foods and consequent HBC seems to increase when natural bear food sources become scarce. For instance, human–grizzly bear conflicts increased in British Columbia, Canada in years of low availability of salmon biomass (Artelle et al. 2016) and similar patterns of increased HBC in periods of natural food failures were described for European brown bears (e.g. Molinari et al. 2016), American black bears (e.g. BaruchMordo et al. 2008), Asiatic black bears (e.g. Oka et al. 2004), and sun bears (Wong et al. 2015). On the other hand, years of high mast production can increase bear reproductive rates (Costello et al. 2003), which in turn can lead to increased levels of conflicts in subsequent years, due to a larger number of mothers with cubs and young individuals that look for shelter and food near humans (Elfström et al. 2014c; Obbard et al. 2014). However, an occurrence of problem bears that is unrelated with seasonal food availability has also been reported (Elfström et al. 2014b). Human population size and the impact of human activities on the landscape also influence the occurrence of HBC. Habitat degradation due to agricultural expansion increases the rates of HBC globally (Can et al. 2014). Additionally, the use of bear habitats by humans can increase the risk of human– bear interaction. For example, direct competition for fruits led to higher rates of sloth bear attacks on humans in India (Dutta et al. 2015). Also, recreational activities, such as bear viewing and nature tourism, are increasingly demanded and without proper regulation can be a cause for more human–bear interactions which may end in HBC (Penteriani et al. 2017). On the other hand, socioeconomic changes may lead to modifications in land use and to land abandonment in rural areas, which, in turn, can facilitate the expansion of bears and thus promote conflicts. For example, in the Ecuadorian Andes shifts from crops to cattle, and from a mostly subsistence-oriented

Figure 15.4 Delay in ocean freezing due to global warming is an important extrinsic factor promoting conflicts between people and polar bears. (Photo by Marcus Elfström.) (A black and white version of this figure will appear in some formats. For the color version, please refer to the plate section.)

economy to a more market-oriented and capital-intensive approach, have led to higher rates of cattle predation by Andean bears (Jampel 2016). Also, decrease and aging of the local human population may enhance leaving unattended fruit trees and unharvested crops, which attract bears to villages (Kishimoto 2009; Yamazaki et al. 2009). Various climatic and meteorological events have been associated with the increasing trend of HBC. The most outstanding example is the considerable loss of sea ice and delay in ocean freezing due to global warming, which is forcing polar bears (Figure 15.4) to increasingly use terrestrial habitats and consequently promotes encounters with people (Wilson et al. 2017) and damage caused by polar bears (Towns et al. 2009; see also Chapter 14). At lower latitudes, an increase in the frequency of crop raiding and approaches to people by sun bears in Indonesia and American black bears in New Mexico (USA) was reported during drought periods associated with the El Niño–Southern Oscillation (Zack et al. 2003; Fredriksson 2012). Climate change is predicted to increase the severity of meteorological events, such as late-spring frosts that are associated with natural food failures in temperate regions and higher use of urban areas by American and Asiatic black bears (Honda 2013; Laufenberg et al. 2018). Additionally, climate change can reduce the duration of bear hibernation and consequently prolong the period when human–bear interactions occur (Johnson et al. 2018), as well as altering the geographic range of natural food resources. As a consequence, bears might be displaced from mountainous areas toward lower, more humanized ones, where an increase in conflicts is expected (Penteriani et al. 2019). Finally, bear-damage management greatly influences the occurrence of HBC (Bautista et al. 2017) and good husbandry practices have been the most effective and widespread technique to prevent conflict with bears and other large carnivores (van Eeden et al. 2018).

217

Human–Bear Coexistence

Human Dimensions of Human–Bear Conflicts At a pragmatic level, it is vital to recognize that effectively addressing and solving HBCs requires working with people, communities, and a variety of stakeholders within a given social and ecological context (see also Chapter 20). Understanding the associated values, perceptions, social and cultural norms, and decision-making of people within a given context is important for illuminating the human dimensions of HBC. While regulatory approaches to mitigating undesirable human behaviors (e.g. poor sanitation practices, lack of use of prevention tools, or illegal poaching) are critically important to addressing HBC, collaborative approaches to decision-making offer additional means to build support and improve human behaviors and practices that address HBC.

Values Among Stakeholders are Important Addressing HBC must have scaled levels of support, built upon effective stakeholder decision-making forums, well-designed management plans, effective public communication strategies, and institutionally backed political, legal, and financial support at national and international governmental levels. In many respects, resolving HBC rests upon the support and acceptance of bears by communities of place and communities of interest – where local and broad public values converge (Wilson 2016). This requires attention to be paid to the values and perspectives of people who actually live with bears and those who have strong interests in bears, often urban dwellers whose values may be different from local stakeholders. Finding areas of convergence in values is no easy task considering that strong differences exist in terms of how groups of people or stakeholders believe bears should be managed and conserved. Moreover, in some contexts, there may be deep-rooted social issues, power inequities, historical events, or ethnic and cultural divides that cause social conflicts, resulting in a refusal by people to work collaboratively to address ways to live with bears. Failure to address these underlying sources of conflict may hinder well-intended efforts and may require facilitated processes that can expose, explore, and transform existing relationships so that meaningful HBC management can occur (Madden & McQuinn 2014). Social conflicts and the actual material damage from bears that impact individuals should therefore be understood within unique social and ecological contexts that reflect a keen understanding of political, cultural, historic, and economic conditions.

Problem Definitions Matter Additionally, how different people orient to and define “the problem” of living with bears is an essential aspect of the human dimensions related to addressing HBC. For example, it is common for stakeholders to define the issue of HBC as one of having “too many bears” (Wilson et al. 2013). The perception that bears are too numerous can be complicated – for example,

218

in certain European contexts, younger bears and females with cubs use human settlements as human shields from older male bears (Elfström et al. 2014c; Steyaert et al. 2016). This can make females and younger bears more regularly visible to people. Subsequently, this can cause the perception that the perceived problem is about an increase in the population of bears, when this may not be the case. If these types of perception are not addressed or corrected, then solutions to the perceived problem may be misguided – in this example, a call to reduce the bear population as a solution to HBC. This is just one example of the importance of paying attention to how local people define “the problem.” In many cases, collaborative forums or decisionmaking processes that bring local people together with bear managers, scientists, and other stakeholders are essential for generating a shared understanding of the problem so that shared solutions are generated for effectively addressing HBC (Botetzagias & Kotilda 2018).

Community Responses to Biological Scale The question of whether individuals and groups of people can work collaboratively to understand the problem and find solutions becomes even more imperative due to the large spatial scale of habitat that bears require. In most European landscapes, brown bear home ranges encompass human activities within a mixed mosaic of modified forest, mountain, and agricultural lands. In this context, the use of prevention tools by multiple individuals is necessary to match the biological scale at which bears live. Failure to match community-level responses to the scale of carnivore life history needs will inevitably fall short of resolving conflicts for both people and bears (Wilson et al. 2013). These challenges are further magnified in areas where large carnivore populations encompass multiple countries and where transboundary movements of bears cross multiple management jurisdictions (Bartoń et al. 2019). In these contexts, well-developed institutional partnerships and transboundary-level management agreements are central for addressing HBC across jurisdictional boundaries (Penteriani et al. 2018b; Proctor et al. 2018).

Effectiveness of Conflict Mitigation Measures During thousands of years of coexistence with bears, humans have developed both proactive and reactive methods to prevent or mitigate HBC. While none of them is 100% effective, there are patterns of success and failure that transcend species. First and foremost, proactive methods that prevent conflict are almost always cheaper and more successful over time than reactive measures. Second, as the levels of food conditioning and human habituation increase, the effectiveness of mitigation decreases. Mast failures and water shortages also influence the effectiveness of mitigation efforts, as well as sex, age, and maternal status of individuals, and competition for resources among individuals (Majić Skrbinšek & Krofel 2015).

Human–Bear Conflicts

Preventing Access to Anthropogenic Food and Human Property As highlighted before, anthropogenic food is a well-known primary cause of HBC; therefore, denying bears access to human food resources is perhaps the most effective and proactive way of preventing HBC. In this way, bears are not rewarded for approaching humans or developed areas, thus habituation to human presence and food conditioning are less likely to occur (Majić Skrbinšek & Krofel 2015). There are numerous approaches to effectively prevent bears from accessing anthropogenic food sources, including livestock, although reliable evidence for effectiveness of many of these methods is still lacking (van Eeden et al. 2018). Exclusion by means of bear-proof bins and electric fences prevents access to unprotected resources such as garbage and livestock, respectively (Sowka 2009). Properly installed and maintained electric fences (Figure 15.5) can have near 100% effectiveness rate and the use of guarding dogs as protection, particularly for livestock, is also a highly effective technique (Belant et al. 2011; Proctor et al. 2018; van Eeden et al. 2018). Removal of attractants such as bird feeders, pet foods, and carcasses is essential, especially in spring or during mast failures, when natural foods are not widely available. Strict legislation and its enforcement, accompanied with public education, are also crucial elements in reducing the availability of anthropogenic foods (Majić Skrbinšek & Krofel 2015).

Aversive Conditioning Aversive conditioning is a learning process in which deterrents are continually and consistently administered to change an undesirable behavior. Negative stimuli, such as shooting with rubber bullets, are applied while an animal is engaged in undesirable behavior in order to elicit an avoidance of such behavior in the future (Gillin et al. 1994). Other methods of aversive conditioning include: taste aversion, cracker shells, propane cannons, bear dogs, shooting with bean bag rounds, and capture and release of conflictual individuals. Most of these methods have produced mixed results and depend on multiple factors, including the context in which a learning process took place, the immediacy of a consequence to a given behavioral response, and the consistency and magnitude of these consequences and rewarding of alternative behavior (Majić Skrbinšek & Krofel 2015). Pain stimuli and the use of bear dogs have proven to be the most successful in the long term. However, effectiveness is lower when an undesired behavior is already strongly established, and it is important to remember that bears may also habituate to some of the aversion techniques (Mazur 2010). A monitoring and response system that quickly detects undesired behaviors is crucial for successful application of this tool.

Lethal Removal Lethal removal of bears has been a widespread measure used in response to bear incidents in the past (Schwartz et al. 2005).

This method includes shooting, trapping, and poisoning. It can refer to management removals of specific “problem” individuals or a general culling of the population with the aim to reduce bear density. In many regions of the world, however, lethal controls are losing public approval, and for endangered populations even limited removal can have strong negative effects (Treves & Karanth 2003; Schwartz et al. 2005). Poisoning and some types of trapping are in general ill-advised due to risk of non-target kills, including endangered wildlife and domestic animals. Lethal methods are most effective when known “problem” bears are removed (Gunther et al. 2004), for which permits may be obtained from a regulatory agency. Culling as a tool for conflict mitigation is not likely an acceptable solution near settlements; however, it may be considered for rural locations with agricultural conflicts, but must also consider population and damage characteristics (Belant et al. 2011).

Non-Lethal Removal This method includes capturing an individual and moving it to another place. Translocations of problem bears are generally more acceptable for the public than lethal removals, but agencies are increasingly reluctant to use this method (Creachbaum et al. 1998). Translocations are costly and labor-intensive with high mortality rates, as bears often return to the capture site even from several hundreds of kilometers away, present road hazards, or cause problems in the new area (Linnell et al. 1997). Noteworthy successes from translocation have considered the age and sex of the individual in addition to significant translocation distance, impassable geographical barriers, and perhaps the quality of the new habitat (Taylor & Phillips 2019). Translocating young males that are not highly foodconditioned has seen the greatest success (Belant et al. 2011; Taylor & Phillips 2019).

Diversionary Feeding of Bears By providing food in remote areas, managers attempt to divert bears from approaching settlements and/or reduce damage to human property. Effectiveness of diversionary feeding lacks rigorous studies and existing examples from brown bears and American black bears have been met with mixed results (Steyaert et al. 2014; Garshelis et al. 2017). Diversionary feeding with carcasses to reduce livestock depredation was not effective (Kavčič et al. 2013; Morehouse & Boyce 2017). There is also an increasing list of side effects, many of which are unwanted, associated with the artificial feeding of bears (Penteriani et al. 2018a). If diversionary feeding is to be used in HBC mitigation, a comprehensive review of the literature suggests that the highest efficacy with minimal side effects is through temporary, seasonal, and as-needed placement of natural foods (Taylor & Phillips 2019). Employing this strategy encourages exploitation of natural food sources first and discourages dependence on anthropogenic supplementary foods.

219

Human–Bear Coexistence

Land-Use Practices There are several potential mechanisms by which land-use practices can affect probability for occurrence of HBC. For example, limiting human access to the most crucial bear habitats in certain time periods gave positive results in American national parks (Coleman et al. 2013). Because cover is an important parameter affecting space use by bears, some authors recommended removing dense vegetation near human settlements (Elfström et al. 2014c) and crops (Sato et al. 2005). Water courses are popular travel routes for bears; therefore, removing livestock and apiaries from these areas can also be effective (Clark et al. 2005). Transition from sheep to cattle or horse breeding, from livestock breeding to other land use (e.g. agriculture, forestry), or selection of crops less attractive to bears can also reduce probability of bear damage (Gunther et al. 2004; Swenson & Andrén 2005; Wilson

et al. 2006; Majić Skrbinšek & Krofel 2015). However, generating a will for such changes among stakeholders can be a considerable challenge (Linnell et al. 2013).

Compensation Economic compensation for claimed damage caused by bears started in the 1950s and is nowadays widespread in Europe and North America (Ravenelle & Nyhus 2017). Compensation programs differ from one country to another and can effectively reduce the economic impact of the HBC on certain stakeholders (Figure 15.6). However, they should be treated with care, as they can also decrease efforts to prevent damage and exacerbate conflicts (Bulte & Rondeau 2005). For example, in the Scandinavian brown bear population the level of conflict is considerably higher in Norway compared to Sweden, which is connected with sheep in Norway being mostly free-ranging and unprotected. In Norway, compensation is given in all cases, even if the incident is not verified and if no protection of sheep was used, whereas in Sweden compensation is conditional on the proper protection of livestock, and wildlife agencies strongly focus on subsidizing preventive measures (Swenson & Andrén 2005; Widman & Elofsson 2018). It should also be understood that compensation does not directly affect the occurrence of HBC, but only mitigates their economic impact.

Public Awareness and Education

Figure 15.5 Electric fences and electric nets, when used properly, are a very effective way to prevent damage caused by bears to livestock, beehives, crops, or other human property. (Photo by Miha Krofel.)

Increasing public awareness about drivers of the HBC, demonstrating the preventive measures and their proper use, can considerably improve the effectiveness of HBC prevention programs. People also tend to follow prescribed rules more when they understand the reasons behind them (Majić Skrbinšek & Krofel 2015). Education of adults and children can also be effective to improve their attitude towards bears (Ambarlı 2016). Figure 15.6 Economic compensation for claimed damages can effectively reduce economic impact of human–bear conflicts, as in the case of damage to fruit trees. Here is shown a brown bear scat full of domestic cherry stones. (Photo by Vincenzo Penteriani.)

220

Human–Bear Conflicts

Successful Examples of Conflict Prevention North American National Parks

reported for the American black bears in residential areas across the USA (Tavss 2005; Leigh & Chamberlain 2008).

The first systematic measures targeting availability of human food sources were applied in the 1970s and 1980s in North American national parks following high rates of HBCs, including several human casualties (Herrero 1994). Strict garbage management, regulations on human food storage, prohibition of artificial bear feeding, and intensive public education about proper behavior in bear habitat proved very successful. After application of these measures, HBC throughout national parks decreased considerably. For example, in the Yellowstone National Park, attacks on people decreased by almost 90% and at the same time there was less need for management removals of bears (Meagher & Phillips 1983; Gunther & Hoekstra 1998). In the Denali National park, cases of bears feeding on anthropogenic food decreased by 96%, which was followed by a 77% decrease in reported HBC and 77% lower number of management removals (Schirokauer & Boyd 1998). Similarly, after the change of focus from bear management to management of people and anthropogenic food, the number of problem bears removed decreased by 94% for black and 86% for brown bears in the Jasper National Park (Ralf 1995), and by 75% for black and 70% for brown bears in the Glacier National Park (Gniadek & Kendall 1998), respectively. In the Yosemite National Park, after management was changed from reactive (lethal removals, translocations, aversive conditioning) to proactive (limiting access to anthropogenic food, education, law enforcement), the proportion of anthropogenic food and garbage in the black bear diet was considerably reduced, which was followed by a 31% decrease in the number of bear incidents and a 63% decrease in the amount of damage caused by bears (Madison 2008; Hopkins et al. 2014).

South America

Residential Areas in North America Limiting the availability of anthropogenic food for bears is generally more difficult in residential areas compared to national parks. However, also here considerable improvements can be achieved with public education and implementation of methods preventing access to anthropogenic food. In Blackfoot Valley (Montana, USA), a proactive, community-based project was launched to provide cost-free removal of livestock carcasses for ranchers; introduce bear-proof garbage bins, provide electric fences for beehives, cattle calving areas, and garbage dumps in parallel with intensive public education. During the three-year project the number of conflicts with brown bears decreased by 91% without the need to remove a single bear (Wilson et al. 2006; Wilson 2007). Substantial decrease in HBC and management removals of brown bears was noted also in Kennecott Valley (Alaska, USA) after local residents were provided with bear-proof garbage containers, electric fences, and targeted public education (Wilder et al. 2007). Similar successes (40–80% reduction) in reducing HBC by preventing bears feeding on human food sources were

To reduce the impact of cattle depredation and damage to corn and avocado plantations caused by Andean bears, a pilot program was developed in Ecuador to improve livestock management by providing water channels, electric fences, fertilizers, grass seeds, artificial insemination, and permanent veterinary care. This resulted in the prevention of further cattle losses and improved farmers’ attitude toward bears and their conservation in the project areas (Laguna 2018). In parallel, bear-watching ecotourism is promoted, which provides additional local employment.

Middle East Electric fencing for preventing brown bear damage to beehives and orchards in Turkey was first implemented at six locations in 2007 and 2008 (Ambarlı & Bilgin 2008). After successful prevention of damage in this pilot EU Kaçkar Mountains Conservation Project and adoption by the government authorities, electric fencing became widespread throughout the country. Today, it is locally manufactured and used at more than 5000 locations, which caused a decrease in bear-caused damage, as well as a 10-fold reduction in the price of the electric fencing equipment compared to imported fences.

Arctic In order to prevent conflicts due to polar bears approaching human settlements, a polar bear alert program has been established in Churchill, Manitoba (Canada), since 1980. Polar bears (with the exception of family groups) that approach residential or working areas of Churchill are captured and after a minimum of 30 days released outside of the town (Hedman 2009). Recaptured individuals are translocated about 70 km north to the Hudson Bay. If translocated bears are recaptured, they may be kept in the holding facility until the ice comes in. The program relies on maintaining a high alert on bear occurrence around the urban area, as well as on educating people to reduce bear accessibility to human-derived foods and restricting human activity outside of the city to avoid disturbing bears. The ice cover is the main driver behind polar bear occurrence ashore and beyond the reach of bear management. However, the bear alert management program results in saving lives of bears and people.

Conclusions Human–bear conflicts are complex and diverse. Consequently, there is no single one-for-all solution to effectively prevent all types of problems. Because often only few problem bears cause the majority of all bear incidents, special attention needs to be given to preventing the development of repetitive conflict behavior, especially preventing bears’ access to anthropogenic food. However, when this fails and bears develop high levels of

221

Human–Bear Coexistence

human habituation or food conditioning, conflict behavior is difficult to change. In such cases, the removal of problem bears is often the only effective solution and will therefore likely remain an important part of bear management in the foreseeable future. Nevertheless, the main focus in bear management should be prevention, as successful proactive management is considerably more acceptable to the public than reactive responses once the conflicts have already occurred. Experience from several regions suggests that this approach gives the best results when local inhabitants are actively involved and

References Ahmad, S., Hameed, S., Ali, H., et al. (2016). Carnivores’ diversity and conflicts with humans in Musk Deer National Park, Azad Jammu and Kashmir, Pakistan. European Journal of Wildlife Research 62: 565–576. Ali, A., Waseem, M., Teng, M., et al. (2018). Human–Asiatic black bear (Ursus thibetanus) interactions in the Kaghan Valley, Pakistan. Ethology Ecology & Evolution 30: 399–415. Ambarlı, H. (2016). Rural and urban students’ perceptions of and attitudes toward brown bears in Turkey. Anthrozoös 29: 489–502. Ambarlı, H. & Bilgin, C.C. (2008). Humanbrown bear conflicts in Artvin, northeastern Turkey: encounters, damage, and attitudes. Ursus 19: 146–153. Artelle, K. A., Anderson, S. C., Reynolds, J. D., et al. (2016). Ecology of conflict: marine food supply affects human– wildlife interactions on land. Scientific Reports 6: 25936. Bargali, H. S., Akhtar, N. & Chauhan, N. P. S. (2005). Characteristics of sloth bear attacks and human casualties in North Bilaspur forest division, Chhattisgarh, India. Ursus 16: 263–267. Bartoń, K. A., Zwijacz-Kozica, T., Zięba, F., Sergiel, A. & Selva, N. (2019). Bears without borders: long-distance movement in human-dominated landscapes. Global Ecology and Conservation 17: p.e00541. Baruch-Mordo, S., Breck, S. W., Wilson, K. R. & Theobald, D. M. (2008). Spatiotemporal distribution of black bear–human conflicts in Colorado, USA. Journal of Wildlife Management 72: 1853–1862. Bautista, C., Naves, J., Revilla, E., et al. (2017). Patterns and correlates of claims for brown bear damage on a continental

222

well-informed about bear biology, occurrence, and prevention of HBC. At the same time, such non-lethal management is in line with recent trends of a decrease in the public’s tolerance of lethal bear removals. The science of HBC has enabled the development of several highly successful conflict-prevention programs throughout the world and has enabled the persistence of numerous bear populations in coexistence with people. Building upon this knowledge and best-practice examples is probably the safest way to ensure the future survival of these iconic animals.

scale. Journal of Applied Ecology 54: 282–292. Beckmann, J. P. & Berger, J. (2003). Rapid ecological and behavioural changes in carnivores: the responses of black bears (Ursus americanus) to altered food. Journal of Zoology 261: 207–212. Bejder, L., Samuels, A., Whitehead, H., Finn, H. & Allen, S. (2009). Impact assessment research: use and misuse of habituation, sensitisation and tolerance in describing wildlife responses to anthropogenic stimuli. Marine Ecology – Progress Series 395: 177–185. Belant, J. L., Simek, S. L. & West, B. C. (2011). Managing human–black bear conflicts. Human–Wildlife Conflicts Monograph 1: 1–77. Borbón-García, A., Reyes, A., Vives-Flórez, M. & Caballero, S. (2017). Captivity shapes the gut microbiota of Andean bears: insights into health surveillance. Frontiers in Microbiology 8: 1–12. Botetzagias, I. & Kotilda, E. (2018). Bear management in Greece as a ‘problem orientation’ policy issue: a Qmethodology approach. Aegean Journal of Environmental Sciences 3: 1–36. Breck, S. W., Williams, C. L., Beckmann, J. P., et al. (2008). Using genetic relatedness to investigate the development of conflict behavior in black bears. Journal of Mammalogy 89: 428–434. Breck, S. W., Lance, N. & Seher, V. (2009). Selective foraging for anthropogenic resources by black bears: minivans in Yosemite National Park. Journal of Mammalogy 90: 1041–1044. Bulte, E. H. & Rondeau, D. (2005). Why compensating wildlife damages may be bad for conservation. Journal of Wildlife Management 69: 14–19. Can, Ö. E., Cruze, N. D., Garshelis, D. L., Beecham, J. J. & Macdonald, D. W. (2014). Resolving human–bear conflict: a global survey of countries, experts and

key factors. Conservation Letters 7: 501–513. Charoo, S. A., Sharma, L. K. & Sathyakumar, S. (2011). Asiatic black bear–human interactions around Dachigam National Park, Kashmir, India. Ursus 22: 106–113. Chauhan, N. P. S. (2003). Human casualties and livestock depredation by black and brown bears in the Indian Himalaya, 1989–98. Ursus 14: 84–87. Clark, J. D., Dobey, S., Masters, D. V., et al. (2005). American black bears and bee yard depredation at Okefenokee Swamp, Georgia. Ursus 16: 234–244. Coleman, T. H., Schwartz, C. C., Gunther, K. A. & Creel, S. (2013). Grizzly bear and human interaction in Yellowstone National Park: an evaluation of bear management areas. The Journal of Wildlife Management 77: 1311–1320. Costello, C. M., Jones, D. E., Inman, R. M., et al. (2003). Relationship of variable mast production to American black bear reproductive parameters in New Mexico. Ursus 14: 1–16. Cozzi, G., Chynoweth, M., Kusak, J., et al. (2016). Anthropogenic food resources foster the coexistence of distinct life history strategies: year-round sedentary and migratory brown bears. Journal of Zoology 300: 142–150. Creachbaum, M. S., Johnson, C. & Schmidt, R. H. (1998). Living on the edge: a process for redesigning campgrounds in grizzly bear habitat. Landscape Urban Plan 42: 269–286. Debata, S., Swain, K. K., Sahu, H. K. & Palei, H. S. (2017). Human–sloth bear conflict in a human-dominated landscape of northern Odisha, India. Ursus 27: 90–98. DeMaster, D. P. & Stirling, I. (1981). Ursus maritimus. Mammalian Species 145: 1–7. Ditmer, M. A., Burk, T. E. & Garshelis, D. L. (2015). Do innate food preferences and learning affect crop raiding by American black bears? Ursus 26: 40–52.

Human–Bear Conflicts

Dutta, T., Sharma, S., Maldonado, J. E., Panwar, H. S. & Seidensticker, J. (2015). Genetic variation, structure, and gene flow in a sloth bear (Melursus ursinus) meta-population in the Satpura-Maikal landscape of Central India. PLoS ONE 10: 7–9. Dyck, M. G. (2006). Characteristics of polar bears killed in defense of life and property in Nunavut, Canada, 1970–2000. Ursus 17: 52–62. Elfström, M., Davey, M. L., Zedrosser, A., et al. (2014a). Do Scandinavian brown bears approach settlements to obtain high-quality food? Biological Conservation 178: 128–135. Elfström, M., Zedrosser, A., Jerina, K., et al. (2014b). Does despotic behavior or food search explain the occurrence of problem brown bears in Europe? Journal of Wildlife Management 78: 881–893. Elfström, M., Zedrosser, A., Stoen, O. G. & Swenson, J. E. (2014c). Ultimate and proximate mechanisms underlying the occurrence of bears close to human settlements: review and management implications. Mammal Review 44: 5–18. Espinosa, S. & Jacobson, S. K. (2012). Human–wildlife conflict and environmental education: evaluating a community program to protect the andean bear in Ecuador. Journal of Environmental Education 43: 55–65. Fredriksson, G. (2005). Human–sun bear conflicts in East Kalimantan, Indonesian Borneo. Ursus 16: 130–137. Fredriksson, G. (2012). Effects of El Nino and large-scale forest fires on the ecology and conservation of Malayan sun bears (Helarctos malayanus) in East Kalimantan, Indonesian Borneo. PhD doctoral thesis, University of Amsterdam. Garcia, K. C., Joshi, H. M. & Dharaiya, N. (2016). Assessment of human–sloth bear conflicts in North Gujarat, India. Ursus 27: 5–10. Garshelis, D. L. (2009). Family Ursidae. In: Wilson, D. E. & Mittermeier, R. A. (Eds.), Handbook of the mammals of the world. Vol. 1. Carnivores (pp. 448–497). Barcelona: Lynx Edicions. Garshelis, D. L., Baruch-Mordo, S., Bryant, A., Gunther, K. A. & Jerina, K. (2017). Is diversionary feeding an effective tool for reducing human–bear conflicts? Case studies from North America and Europe. Ursus 28: 31–55. Ghadirian, T., Qashqaei, A. T., Soofi, M., Abolghasemi, H. & Ghoddousi, A. (2017).

Diet of Asiatic black bear in its westernmost distribution range, southern Iran. Ursus 28: 15–19. Gillin, C. M., Hammond, F. M. & Peterson, C. G. (1994). Evaluation of an aversive conditioning technique used on female grizzly bears in the Yellowstone ecosystem. International Conference on Bear Research and Management 9: 503–512. Gniadek, S. J. & Kendall, K. C. (1998). A summary of bear management in Glacier National Park, Montana, 1960–1994. Ursus 10: 155–159. Goldstein, I. (1991). Spectacled bear predation and feeding behavior on livestock in Venezuela. Studies on Neotropical Fauna and Environment 26: 231–235. Groff, C., Dalpiaz, D., Frapporti, C., Rizzoli, R. & Zanghellini, P. (2010). 2009 Bear report of the Forestry and Wildlife Department of the Autonomous Province of Trento. Trento: Autonomous Province of Trento. Guharajan, R., Abram, N. K., Magguna, M. A., et al. (2017). Does the vulnerable sun bear Helarctos malayanus damage crops and threaten people in oil palm plantations? Oryx 134: 635–646. Gunther, K. A. & Hoekstra, H. E. (1998). Bear-inflicted human injuries in Yellowstone National Park, 1970–1994. Ursus 10: 377–384. Gunther, K. A., Haroldson, M. A., Frey, K., et al. (2004). Grizzly bear–human conflicts in the Greater Yellowstone ecosystem, 1992–2000. Ursus 15: 10–22. Hedman, D. (2009). Manitoba Conservation, Thompson, Canada. Summary of the third international bear–people conflicts workshop in Canmore, Alberta (November 15–18), and oral presentation. Herrero, S. (1994). The Canadian National Parks and grizzly bear ecosystems: the need for interagency management. International Conference on Bear Research and Management 9: 7–21. Herrero, S., Smith, T., Debruyn, T. D., Gunther, K. & Matt, C. A. (2005). From the field: brown bear habituation to people – safety, risks, and benefits. Wildlife Society Bulletin 33: 362–373. Honda, T. (2013). Late spring frosts induce human–Asiatic black bear conflicts. Mammal Study 38: 287–292.

Hopkins, J. B. I., Koch, P. L., Ferguson, J. M. & Kalinowski, S. T. (2014). The changing anthropogenic diets of American black bears over the past century in Yosemite National Park. Frontiers in Ecology and the Environment 12: 107–114. Huang, C., Li, X.-Y., Shi, L.-J. & Jiang, X.-L. (2018). Patterns of human–wildlife conflict and compensation practices around Daxueshan Nature Reserve, China. Zoological Research 39: 406–412. Immelmann, K. & Beer, C. (1989). A dictionary of ethology. Cambridge, MA: Harvard University Press. Jampel, C. (2016). Cattle-based livelihoods, changes in the taskscape, and human– bear conflict in the Ecuadorian Andes. Geoforum 69: 84–93. Jamtsho, Y. & Wangchuk, S. (2016). Assessing patterns of human–Asiatic black bear interaction in and around Wangchuck Centennial National Park, Bhutan. Global Ecology and Conservation 8: 183–189. Johnson, H. E., Lewis, D. L., Verzuh, T. L., et al. (2018). Human development and climate affect hibernation in a large carnivore with implications for human– carnivore conflicts. Journal of Applied Ecology 55: 663–672. Jonker, S. A., Parehurst, J. A., Field, R. & Fuller, T. K. (1998). Black bear depredation on agricultural commodities in Massachusetts. Wildlife Society Bulletin 26: 318–324. Jorgenson, J. P. & Sandoval, A. S. (2005). Andean bear management needs and interactions with humans in Colombia. Ursus 16: 108–116. Kaczensky, P. (1999). Large carnivore depredation on livestock in Europe. Ursus 11: 59–72. Karamanlidis, A. A., Sanopoulos, A., Georgiadis, L. & Zedrosser, A. (2011). Structural and economic aspects of human–bear conflicts in Greece. Ursus 22: 141–151. Kavčič, I., Adamič, M., Kaczensky, P., Krofel, M. & Jerina, K. (2013). Supplemental feeding with carrion is not reducing brown bear depredations on sheep in Slovenia. Ursus 24: 111–119. Kimball, B. A., Nolte, D. L., Engeman, R. M., Johnston, J. J. & Stermitz, F. R. (1998). Chemically mediated foraging preference of black bears (Ursus americanus). Journal of Mammalogy 79: 448–456.

223

Human–Bear Coexistence

Kishimoto, R. (2009). Status of the 2006 drastic rise of Asiatic black bear (Ursus thibetanus) intrusions into residential areas in Nagano Prefecture. In: Oi, T., Ohnishi, N., Koizumi, T. & Okochi, I. (Eds.), Biology of bear intrusions. Proceedings of the International Workshop “The Mechanism of the Intrusion of Bears into Residential Areas,” November 21–22, 2008, Kyoto, Japan (pp. 35–39). Ibaraki: Forestry and Forest Products Research Institute. Kobashikawa, S. & Koike, S. (2016). Spatiotemporal factors affecting bark stripping of conifer trees by Asiatic black bears (Ursus thibetanus) in Japan. Forest Ecology and Management 380: 100–106. Krofel, M. & Jerina, K. (2012). Review of human–bear conflicts: causes and possible solutions. Professional Journal of Forestry 70: 235–253. Krofel, M., Špacapan, M. & Jerina, K. (2017). Winter sleep with room service: denning behaviour of brown bears with access to anthropogenic food. Journal of Zoology 302: 8–14 Laguna, A. (2018). Manual de prevención y atención de ataques por Oso Andino (Tremarctos ornatus) al ganado en Ecuador. Quito: Big Mammals Conservation.

Lira Torres, I., Carrera Treviño, R. & Martínez García, L.F. (2015). Primer registro de oso negro Ursus americanus (Carnivora: Ursidae) en un bosque tropical subperennifolio en la reserva de la biosfera “El Cielo”, Tamaulipas, México. Therya 6: 653–660. Liu, F., McShea, W. J., Garshelis, D. L., et al. (2011). Human–wildlife conflicts influence attitudes but not necessarily behaviors: factors driving the poaching of bears in China. Biological Conservation 144: 538–547. Lowery, D. R., Morse, W. C. & Steury, T. D. (2012). Biological and social investigation of human–black bear conflicts in the panhandle of Florida. Human Dimensions of Wildlife 17: 193–206. Madden, F. & McQuinn, B. (2014). Conservation’s blind spot: the case for conflict transformation in wildlife conservation. Biological Conservation 178: 97–106. Madison, J. S. (2008). Yosemite National Park: the continuous evolution of human–black bear conflict management. Human–Wildlife Conflicts 2: 160–167. Majić Skrbinšek, A. & Krofel, M. (2015). Defining, preventing, and reacting to problem bear behaviour in Europe (pp. 1–56). Brussels: European Commission.

Laufenberg, J. S., Johnson, H. E., Doherty, P. F. & Breck, S. W. (2018). Compounding effects of human development and a natural food shortage on a black bear population along a human development– wildland interface. Biological Conservation 224: 188–198.

Mattson, D. J., Blanchard, B. M. & Knight, R. R. (1992). Yellowstone grizzly bear mortality, human habituation, and whitebark pine seed crops. Journal of Wildlife Management 56: 432–442.

Leigh, J. & Chamberlain, M. J. (2008). Effects of aversive conditioning on behavior of nuisance Louisiana black bears. Human– Wildlife Conflicts 2: 175–182.

Mazur, R. L. (2010). Does aversive conditioning reduce human–black bear conflict? The Journal of Wildlife Management 74: 48–54.

Li, X., Buzzard, P., Chen, Y. & Jiang, X. (2013). Patterns of livestock predation by carnivores: human–wildlife conflict in Northwest Yunnan, China. Environmental Management 52: 1334–1340.

Mazur, R. & Seher, V. (2008). Socially learned foraging behaviour in wild black bears, Ursus americanus. Animal Behaviour 75: 1503–1508.

Linnell, J. D. C., Aanes, R., Swenson, J. E., Odden, J. & Smith, M. E. (1997). Translocation of carnivores as a method for managing problem animals: a review. Biodiversity and Conservation 6: 1245–1257. Linnell, J. D. C., Lescureux, N., Majic, A., Von Arx, M. & Salvatori, V. (2013). Results from a stakeholder workshop on

224

large carnivores in Brussels (pp. 1–68). Brussels: European Commission.

McCullough, D. R. (1982). Behavior, bears, and humans. Wildlife Society Bulletin 10: 27–33. McKinley, B. K., Belant, J. L. & Etter, D. R. (2014). American black bear–apiary conflicts in Michigan. Human–Wildlife Interactions 8: 228–234. McLellan, B. N. (2005). Sexually selected infanticide in grizzly bears: the effects of hunting on cub survival. Ursus 16: 141–156.

Meagher, M. & Phillips, J. R. (1983). Restoration of natural populations of grizzly and black bears in Yellowstone National Park. International Association of Bear Research and Management 5: 152–158. Merkle, J. A., Robinson, H. S., Krausman, P. R. & Alaback, P. (2013). Food availability and foraging near human developments by black bears. Journal of Mammalogy 94: 378–385. Miller, S. D. & Tutterrow, V. L. (1999). Characteristics of nonsport mortalities to brown and black bears and human injuries from bears in Alaska. Ursus 11: 239–252. Mir, Z. R., Noor, A., Habib, B. & Veeraswami, G. G. (2015). Attitudes of local people toward wildlife conservation: a case study from the Kashmir Valley. Mountain Research and Development 35: 392–400. Molinari, P., Krofel, M., Bragalanti, N., et al. (2016). Comparison of the occurrence of human–bear conflicts between northern Dinaric Mountains and south-eastern Alps. Carnivore Damage Prevention News 12: 9–17. Morehouse, A. T. & Boyce, M. S. (2017). Evaluation of intercept feeding to reduce livestock depredation by grizzly bears. Ursus 28: 66–80. Morehouse, A. T., Graves, T. A., Mikle, N. & Boyce, M. S. (2016). Nature vs. nurture: evidence for social learning of conflict behaviour in grizzly bears. PLoS ONE 11: e0165425. Mueller, C., Herrero, S. & Gibeau, M. L. (2004). Distribution of subadult grizzly bears in relation to human development in the Bow River Watershed, Alberta. Ursus 15: 35–47. Narita, R., Mano, T., Yokoyama, R. & Takayanagi, A. (2011). Variation in maize consumption by brown bears (Ursus arctos) in two coastal areas of Hokkaido, Japan. Mammal Study 36: 33–39. Naves J., Ordiz A., Fernández-Gil A., et al. (2018) Patterns of brown bear damages on apiaries and management recommendations in the Cantabrian Mountains, Spain. PLoS ONE 13: e0206733. Normua, F., Higashi, S., Ambu, L. & Mohamed, M. (2004). Notes on oil palm plantation use and seasonal spatial relationships of sun bears in Sabah, Malaysia. Ursus 15: 227–231.

Human–Bear Conflicts

Northrup, J. M., Stenhouse, G. B. & Boyce, M. S. (2012). Agricultural lands as ecological traps for grizzly bears. Animal Conservation 15: 369–377. Obbard, M. E., Howe, E. J., Wall, L. L., et al. (2014). Relationships among food availability, harvest, and human–bear conflict at landscape scales in Ontario, Canada. Ursus 25: 98–110. Oka, T., Miura, S., Masaki, T., et al. (2004). Relationship between changes in beechnut production and Asiatic black bears in northern Japan. Journal of Wildlife Management 68: 979–986. Penteriani, V., López-Bao, J. V., Bettega, C., et al. (2017). Consequences of brown bear viewing tourism: a review. Biological Conservation 206: 169–180. Penteriani, V., Delgado, M. D. M., Krofel, M., et al. (2018a). Evolutionary and ecological traps for brown bears Ursus arctos in human-modified landscapes. Mammal Review 48: 180–193. Penteriani, V., Huber, D., Jerina, K., et al. (2018b). Trans-boundary and transregional management of a large carnivore: managing brown bears across national and regional borders in Europe. In: Hovardas, T. (Ed.), Large carnivore conservation and management: Human dimensions and governance. Oxford, UK: Routledge, Taylor & Francis Group. Penteriani V., Zarzo-Arias A., NovoFernández, A., Bombieri G. & LópezSánchez, C. A. (2019). Responses of an endangered brown bear population to climate change based on predictable food resource and shelter alterations. Global Change Biology 25: 1133–1151. Proctor, M. F., Kasworm, W. F., Annis, K. M., et al. (2018). Conservation of threatened Canada–USA trans-border grizzly bears linked to comprehensive conflict reduction. Human–Wildlife Interactions 12: 348–372. Rajpurohit, K. S. & Krausman, P. R. (2000). Human–sloth bear conflicts in Madhya Pradesh, India. Wildlife Society Bulletin 28: 393–399. Ralf, R. (1995). History of bear–human conflict management in Jasper National Park: 1907–1995. Jasper: Jasper National Park. Ravenelle, J. & Nyhus, P. J. (2017). Global patterns and trends in human–wildlife conflict compensation. Conservation Biology 31: 1247–1256. Saito, M., Yamauchi, K. & Aoi, T. (2008). Individual identification of Asiatic black

bears using extracted DNA from damaged crops. Ursus 19: 162–167. Sangay, T. & Vernes, K. (2008). Human– wildlife conflict in the Kingdom of Bhutan: patterns of livestock predation by large mammalian carnivores. Biological Conservation 141: 1272–1282. Sato, Y., Mano, T. & Takatsuki, S. (2005). Stomach contents of brown bears Ursus arctos in Hokkaido, Japan. Wildlife Biology 11: 133–144. Schirokauer, D. W. & Boyd, H. M. (1998). Bear–human conflict management in Denali National Park and Preserve, 1982–94. Ursus 10: 395–403. Schwartz, C. C., Swenson, J. E. & Miller, S. D. (2005). Large carnivores, moose, and humans: a changing paradigm of predator management in the 21st century. Alces 39: 41–63. Schwartz, C. C., Haroldson, M. A., White, G. C., et al. (2006). Temporal, spatial, and environmental influences on the demographics of grizzly bears in the Greater Yellowstone Ecosystem. Wildlife Monographs 161: 1–68. Scotson, L., Vannachomchan, K. & Sharp, T. (2014). More valuable dead than deterred? Crop-raiding bears in Lao PDR. Wildlife Society Bulletin 38: 783–790. Sethy, J. & Chauhan, N. S. (2013). Human– sun bears conflict in Mizoram, North East India: impact and conservation management. International Journal of Conservation Science 4: 317–328. Shivik, J. A., Ruid, D., Willging, R. C. & Mock, K. E. (2011). Are the same bears repeatedly translocated from corn crops in Wisconsin? Ursus 20: 114–119. Sowka, P. (2009). Techniques and refuse management options for residential areas, campgrounds, and group-use area. In: Living with Black Bears, Grizzly Bears and Lions Project. 3. Swan Valley, MT: Living with Wildlife Foundation, & Montana Fish, Wildlife and Parks. Stenhouse, G. B., Lee, L. J. & Poole, K. G. (1988). Some characteristics of polar bear killed during conflicts with humans in the Northwest Territories, 1976–86. Arctic 41: 275–278. Stewart, W. B., Witmer, G. W. & Koehler, G. M. (1999). Black bear damage to forest stands in western Washington. Western Journal of Applied Forestry 14: 128–131. Steyaert, S. M. J. G., Kindberg, J., Swenson, J. E. & Zedrosser, A. (2013). Male reproductive strategy explains

spatiotemporal segregation in brown bears. Journal of Animal Ecology 82: 836–845. Steyaert, S. M., Kindberg, J., Jerina, K., et al. (2014). Behavioral correlates of supplementary feeding of wildlife: can general conclusions be drawn? Basic and Applied Ecology 15: 669–676. Steyaert, S. M. J. G., Leclerc, M., Pelletier, F., et al. (2016). Human shields mediate sexual conflict in a top predator. Proceedings of the Royal Society B 283: 20160906. Swenson, J. E. & Andrén, H. (2005). A tale of two countries: large depredation and compensation schemes in Sweden and Norway. In: Woodroffe, R., Thirgood, S. & Rabinowitz, A. (Eds.), People and wildlife, conflict or coexistence? (pp. 323–339). Cambridge, UK: Cambridge University Press. Swenson, J. E., Sandegren, F., Soderberg, A., et al. (1997). Infanticide caused by hunting of male bears. Nature 386: 450–451. Swenson, J. E., Dahle, B. & Sandegren, F. (2001). Intraspecific predation in Scandinavian brown bears older than cubs-of-the-year. Ursus 12: 81–92. Tavss, E. A. (2005). Correlation of reduction in nuisance black bear complaints with implementation of a nonviolent program and a hunt. In; New Jersey public hearing on the comprehensive black bear management policy. Rutgers, NJ: State University of New Jersey. Taylor, J. D. & Phillips, P. (2019). Black bears. Wildlife Damage Management Series. U.S. Department of Agriculture Animal and Plant Health Inspection Service: Wildlife Services. Towns, L., Derocher, A. E., Stirling, I., Lunn, N. J. & Hedman, D. (2009). Spatial and temporal patterns of problem polar bears in Churchill, Manitoba. Polar Biology 32: 1529–1537. Treves, A. & Karanth, K. U. (2003). Human– carnivore conflict and perspectives on carnivore management worldwide. Conservation Biology 17: 1491–1499. Treves, A., Kapp, K. J. & Macfarland, D. M. (2010). American black bear nuisance complaints and hunter take. Ursus 21: 30–42. van Eeden, L. M., Eklund, A., Miller, J. R., et al. (2018). Carnivore conservation needs evidence-based livestock protection. PLoS Biology 16: e2005577.

225

Human–Bear Coexistence

Widman, M. & Elofsson, K. (2018). Costs of livestock depredation by large carnivores in Sweden 2001 to 2013. Ecological Economics 143: 188–198.

Landscape conditions predisposing grizzly bears to conflicts on private agricultural lands in the western USA. Biological Conservation 130: 47–59.

Wielgus, R. B. & Bunnell, F. L. (1994). Sexual segregation and female grizzly bear avoidance of males. Journal of Wildlife Management 58: 405–413.

Wilson, S. M., Neudecker, G. A. & Jonkel, J. J. (2013). Human–grizzly bear coexistence in the Blackfoot River watershed, Montana: getting ahead of the conflict curve. In: Clark, S. & Rutherford, M. (Eds.), Large carnivore conservation: Integrating science and policy in the North American West (pp. 117–214). Chicago, IL: University of Chicago Press.

Wilder, J. M., DeBruyn, T. D., Smith, T. S. & Southwould, A. (2007). Systematic collection of bear–human interaction information for Alaska’s national parks. Ursus 18: 209–216. Wilson, R. R., Regehr, E. V., St. Martin, M., et al. (2017). Relative influences of climate change and human activity on the onshore distribution of polar bears. Biological Conservation 214: 288–294. Wilson, S. M. (2007). Community-supported conservation of grizzly bears on private agricultural lands. Final close-out report for conservation innovation grant. Portland, OR: US Department of Agriculture – Natural Resources Conservation Service. Wilson, S. M. (2016). A guidebook to human– carnivore conflict: Strategies and tips for effective communication and collaboration with communities (pp. 1–67). Ljubljana, Slovenia: Slovenia Forest Service – LIFE DINALP BEAR Project. Wilson, S. M., Madel, M. J., Mattson, D. J., Graham, J. M. & Merrill, T. (2006).

226

Wong, W. M., Leader-Williams, N. & Linkie, M. (2015). Managing human–sun bear conflict in Sumatran agroforest systems. Human Ecology 43: 255–266. Wynn-Grant, R., Ginsberg, J. R., Lackey, C. W., Sterling, E. J. & Beckmann, J. P. (2018). Risky business: modeling mortality risk near the urban-wildland interface for a large carnivore. Global Ecology and Conservation 16: e00443. Yamazaki, K., Koike, S., Kozakai, C., et al. (2009). Behavioral study of free ranging Japanese black bears I. Does food abundance affect the habitat use of bears? In: Oi, T., Ohnishi, N., Koizumi, T. & Okochi, I. (Eds.), Biology of bear intrusions. Proceedings of the International Workshop “The Mechanism of the Intrusion of Bears into Residential Areas,” November 21–22, 2008, Kyoto, Japan (pp. 60–63). Ibaraki:

Forestry and Forest Products Research Institute. Zack, C. S., Milne, B. T. & Dunn, W. C. (2003). Southern oscillation index as an indicator of encounters between humans and black bears in New Mexico. Wildlife Society Bulletin 31: 517–520. Zapata-Ríos, G. & Branch, L. C. (2018). Mammalian carnivore occupancy is inversely related to presence of domestic dogs in the high Andes of Ecuador. PLoS ONE 13: p.e0192346. Zedrosser, A., Steyaert, S. M. J. G., Gossow, H. & Swenson, J. E. (2011). Brown bear conservation and the ghost of persecution past. Biological Conservation 144: 2163–2170. Ziegltrum, G. J. & Nolte, D. L. (2001). Black bear forest damage in Washington state, USA: economic, ecological, social aspects. Ursus 12: 169–172. Zukowski, B. & Ormsby, A. (2016). Andean bear livestock depredation and community perceptions in Northern Ecuador. Human Dimensions of Wildlife 21: 111–126. Zyśk-Gorczyńska, E., Jakubiec, Z., Wertz, B. & Wuczyński, A. (2016). Long-term study of damage to trees by brown bears Ursus arctos in Poland: increasing trends with insignificant effects on forest management. Forest Ecology and Management 366: 53–64.

Chapter

16

Principles of Human–Bear Conflict Management in Challenging Environments ¨ zgu¨n Emre Can O

The problem of knowledge is that there are many more books on birds written by ornithologists than books on birds written by birds and books on ornithologists written by birds. Nassim Nicholas Taleb

Humans and bears have coexisted since time immemorial. Ötzi the Iceman, a 5300-year-old natural mummy discovered in the Alps on the Italian–Austrian border on September 19, 1991 was found to be wearing a fur hat made from brown bear fur (O’Sullivan et al. 2016). Ötzi’s hat protected his head from the elements, so in other words, brown bears made it possible for Ötzi (who died at the age of 45) and his contemporaries to survive in an extreme environment some 5300 years ago. Today, millions of people live happily alongside bears on four continents. As a result, millions of interactions occur between people and bears. Most of these interactions end without injury to a person, but there are places where problems between people and bears exist (Can et al. 2014). According to the United Nations, in just 30 years or so, 87% of the world’s population, about 7.8 billion people, will be living in developing countries (United Nations 2004). Therefore, in places where bears still exist, more people will find themselves in close proximity to bears, which will likely result in increased problems between people and bears. Failure to mitigate those problems will likely reduce the tolerance of local communities and diminish conservation efforts, not only for bears but also for wider initiatives in general. Therefore, effective problem-solving, or as it is called, human–bear conflict management, will be needed more than ever within the next three decades, but what actually is human–bear conflict (Box 16.1)? An international group of experts at a workshop in Istanbul, Turkey defined human–bear conflict as “any situation where wild bears undesirably use or damage human property, where wild bears harm people, or where people perceive bears to be a direct threat to their property or safety” (WSPA 2009). This “Istanbul definition” of conflict was later used by the Human Bear Conflict Expert Team of IUCN SSC Bear Specialist Group (IUCN SSC BSG 2019).

Box 16.1 On the term “human–bear conflict” and the need for a more accurate one The so-called “conflict” between humans and wildlife has been ongoing since time immemorial. The word “conflict” comes from the Latin word “conflīctus” and means a fight, battle, a prolonged struggle, or clashing or variance of opposed principles, statements, arguments (OED 2015). As bears and other wildlife species are not purposefully seeking to undermine human interests, it is evident that terms such as “human–bear conflict” and “human–wildlife conflict” are more of a metaphor rather than technical terms. Metaphors might be useful and contribute to effective communication. Militarism is a powerful mode of conceptualizing real-world challenges (Larson 2005). However, although military terminology sometimes helps us to grasp daily life concepts rapidly, it can be damaging with unintended consequences as well. For example, Elm and Diener (2007) concluded that military terminology is corrupting biomedical communication and warned physicians and researchers to avoid the cynical logic of war when writing about care of patients. Militaristic metaphors may affect the perceptions of scientific objectivity because of their resonance within contemporary political settings (Larson 2005). Using militaristic language to define the interactions between people and wildlife could weaken the current efforts for nature conservation, and cause loss of scientific credibility (Larson 2005; Redpath et al. 2015). In the case of human–wildlife conflict, the metaphor reinforces the notion that humans are apart from nature; not a part of it (Peterson et al. 2010, 2013). Although metaphors might motivate conservation action in the short term, in the long term they could be ineffective and can even cause misperceptions among conservationists themselves (Larson 2005) and the public. Scientific research has recently shown that the language one uses shapes thinking about the fundamental dimensions of human experience such as space, time, causality, and relationships to others (Boroditsky 2011). Evidence-based science now shows us that language profoundly influences the way we see the world (Boroditsky 2010). The implication of linguistic

227

Human–Bear Coexistence

Box 16.1 (cont.) research for wildlife management and conservation is that the way we frame the interaction between people and wildlife actually does matter. Larson (2005) warns us that linguistic wars can contribute to real wars. As history has proved with the case of people’s long-lasting efforts to exterminate wolves in Europe and North America, once humans declare “war” against another species, the other species does not have much chance to survive. Therefore, those who are using the metaphor human–bear conflict have to consider how this metaphor might be affecting people’s reality. This is because, as the National Academy of Sciences (Institute of Medicine 2009) puts it, “Research is based on the same ethical values that apply in everyday life, including honesty, fairness, objectivity, openness, trustworthiness, and respect for others.” There is a need for a language that is consistent with a vision of human–wildlife coexistence, not one that contributes to polarized and militaristic conceptualizations of management and conservation issues (Larson 2005). Do we have an excuse to ascribe value and attachment to the term human–wildlife conflict merely because it has been increasingly used in the literature in the last decade or so? The term “human–bear conflict” is not problem-free, it frames the interaction between people and bears as negative and we need to resist metaphoric thinking and find accurate descriptions for effective management and conservation. A widely accepted term to replace the word conflict has yet to be found. Therefore, for practical reasons, I use the term human–bear conflict in this chapter to describe the problems that arise between people and bears. I ask our readers not to allow the definitional sensitivity of the word “conflict” to distract from the focus of this chapter.

Human–bear conflicts differ geographically across the range of eight bear species (Can et al. 2014; see also Chapter 15). Although it is difficult to quantify actual trends across the range of bear species, human–bear conflict is perceived to be increasing across the globe (Hristienko & McDonald 2007). Recently, a global survey (Can et al. 2014) involving 104 experts submitting 130 completed questionnaires (as some experts provided information on more than one bear species) found that most experts (89%, n = 115) reported a high level of conflict in their area of expertise; conflicts were increasing in more than half of the areas from which survey respondents responded (56%, n = 73), and the majority (76%, n = 86) of the experts concluded that human–bear conflict has had a negative impact on conservation of bears on all four continents where bears are present (Can et al. 2014). Later, a survey carried out among the members of the IBA also showed that conflicts are considered to be a major issue for the next decade (Zedrosser 2019). In brief, there is consensus among the world’s experts that the problem of human–bear conflict is worsening across the range of bear species. However, the same problem has some quite different impacts on people and bears in Western Educated Industrialized Rich Democratic countries (or WEIRD countries, hereafter) compared to nonWEIRD countries. A whistle-stop tour around the globe

228

indicates that: in North America conflicts do not pose a threat to the viability of American black (Ursus americanus) or brown bears (U. arctos) but cause annoyance for people; whereas in Europe, illegal killing of bears as a retaliation against real or perceived threats is the main cause of bear mortality (Can et al. 2014). On the other hand, conflicts cause hardships for people and affect the rural economy in Asia and South America (Can et al. 2014; see also Chapter 15). In return, retaliation against bears threatens the viability of Asiatic black bears and Andean bears in some of parts of their range and hinders acceptance of conservation initiatives (Can et al. 2014). Conflict is affecting an increasing number of people in Asia and South America (Can et al. 2014). However, conservation practitioners in Asia and South America often operate in challenging circumstances, in many cases without the resources that are available in North America or Europe. By adapting the definition made by Paulus et al. (2009) for extreme environments (such as combat situations, expeditions at high altitudes, or in space), I define a challenging environment for conservation as an “environment that exposes the practitioner to demanding psychological and physical conditions and that requires optimal cognitive and behavioral performance and where institutional structure for wildlife management and conservation is limited or non-existent.” Typically, such areas are in resource-limited countries, but resource-limited islands can also be found in resource-rich countries. It is a challenge to write something useful for practitioners working on human–bear conflict in resource-limited environments in a limited space. Nevertheless, I will make an attempt to do so in this chapter by utilizing: (1) the relevant literature; (2) review of 44 management and action plans from 13 countries (Albania, Austria, Bulgaria, Canada, Croatia, Georgia, Italy, Latvia, Poland, Romania, Slovenia, USA, and Venezuela); (3) the results of the first global survey on human–bear conflicts (Can et al. 2014); and (4) the hard-earned wisdom of a series of participatory workshops that I co-organized (Turkey in 2002, 2008, 2010, and 2012; Greece in 2015; Ecuador in 2017) and was invited to (Georgia in 2010; Nepal 2015; UK in 2016 and 2017). I will first give a brief summary of how conflict studies started in international relationships and then entered the field of wildlife conservation. Then, I will provide an overall picture of human–bear conflict management and discuss how current conflict management plans can be improved by considering them as logic models. Finally, I will present 12 key factors for managing human–bear conflict in challenging environments.

How did the Conflict Resolution Studies Begin? As Swenson in Wilson (2016) stated succinctly, managing human–carnivore conflicts “requires a knowledge of the science behind carnivore biology and human conflict management in addition to the insight that comes only from actual experience.” Therefore, successful human–bear conflict management requires a melding of science (which depends on structured analytical

Principles of Human–Bear Conflict Management

Figure 16.1 Google Books Ngram Viewer graph shows the increase in the use of the terms “conflict” and “conflict resolution” in books published since 1800 and those that are digitized by Google.

methods) and art (which depends on intuitive judgment and personal experience). The “science” and “art” of conflict management has been practiced in North America and Europe for the last seven decades.1 In the beginning, conflict resolution had nothing to do with bears or even with wildlife. A small group of pioneering researchers spotted the need for studying conflict itself in the 1950s after the Cold War, when a potential nuclear war was threatening the future of humanity and conflict resolution became a distinct field of study (Ramsbotham et al. 2011). The term “conflict resolution” was first used to define a new field of study with the first edition of the Journal of Conflict in 1957. Although it was a journal in the field of social sciences, the first editorial article of the Journal of Conflict (Editors 1957) was presenting the new journal surprisingly as a new species with the name Interdisciplinaris internationalis in Linnean terms. Conflict resolution studies progressed over time and were instrumental for making a positive difference in international relations as well as domestic politics in South Africa, the Middle East, Central America, and South-East Asia by the 1980s (Ramsbotham et al. 2011; Figure 16.1). It is not known when conflict resolution was first used in environmental disputes, but the practice became common in the 1980s (Cohn 2002). In his landmark paper, “Tragedy of the Commons,” Hardin (1968) presented how environmental conflicts arouse. Environmental and resource issues are the main reasons for global conflict today (Ramsbotham et al. 2011).

Conflict Resolution in Wildlife Conservation

Presnall (1943) was perhaps the first author who mentioned conflict in relation to wildlife conservation in a scholarly journal. According to him, understanding the conflict concepts of American Indians and Caucasians in relation to wildlife values was vital 1

for effective game conservation programs on Indian lands. Presnall (1943) stated “the basic concepts of two races are not very different, the Indian has merely applied vocation concepts to wildlife and we think of it as an avocation.” Although he was not using the term “conflict resolution,” he was trying to find a common ground for effective game conservation in his article and referring to what we call today “conflict resolution” in management and conservation. Today, when the term conflict resolution is used in relation to wildlife management and conservation, what is actually implied is management of problems that arise between humans and wildlife, as resolution in the literal meaning of the word rarely happens. Even reducing impacts of conflict may not resolve conservation conflicts; in fact, perhaps no conservation conflict has ever been fully resolved by eliminating that conflict (Redpath et al. 2013; Pooley et al. 2017). Today, the phrase “human–wildlife conflict” has almost replaced the term “human–wildlife relationship” and almost any animal behavior can now be described as human–wildlife conflict if that behavior can be perceived as a problem by one or more people (Peterson et al. 2013; Can et al. 2014). Moreover, human–human conflicts are now often said to be a component of human–wildlife conflict. However, a dispute between people – even if it might be related with wildlife or environmental issues – is a dispute between people. Failure to deal with disputes between people only contributes to escalations of such disputes between humans, making such disputes much harder to resolve (Peterson et al. 2013). For example, see Pooley et al. (2017) to learn how rural hunters kill wolves (Canis lupus) to express anger to the state in Sweden.

Human–Bear Conflict Management

Probably the first international gathering about human–bear conflict management was the international human–bear conflict

See Artelle et al. (2018), who report that most management systems in North America lack indications of the basic elements of a scientific approach.

229

Human–Bear Coexistence

Human–Bear Conflict Management

Human-focused methods

Education

Bear-focused methods

Avoiding negative encounters Non-lethal

Lethal

Management of attractants

Compensation & insurance programs

Use of physical barriers

Habitat management

Legislation development & enforcement

Awareness raising

Use of bear deterrents

Diversionary & supplementary feeding

Early warning approaches

Landscape planning

Aversive behavioral conditioning

Relocation & translocation

Bear population management

Traditional methods

Removal of conflict individuals

Figure 16.2 The toolbox of human–bear conflict management consists of 17 tools.

workshop in Yellowknife, Canada in 1987 where few attended (Matt 2012). Since then, human–bear conflict and its management have become a major topic for researchers and practitioners. At the heart of conflict management lies mitigating, compensating, and controlling the damage caused by bears and, in an ideal world, conflict is managed by operational plans called human–bear conflict management plans. These plans are simply a detailed road map to make the world a better place by changing the initial state of conflict to a desired state of conflict in a given area. A re-examination of data from Can et al. (2014) showed that

230

conflict management can be said to include 17 tools that are targeted to humans and bears when 44 management plans totalling about 2375 pages are considered (Figure 16.2). For effective conflict management, these 17 tools need to be used in a tactical way and within the framework of a carefully constructed strategy. However, there are no established standards for preparing conflict plans. Apart from 9% (n = 4), most of the reviewed plans were crowded with text and the actions listed had no specified time frame, making the documents an inventory rather than a management plan. Similarly, only 9%

Principles of Human–Bear Conflict Management

(n = 4) of the plans mentioned the welfare of bears, the elephant-in-the-room aspect of conflict management. On a positive note, it is promising that 36% (n = 16) of the documents were collaborative productions of government agencies, non-governmental organizations (NGOs), and researchers, but it is less encouraging that there was little evidence of engagement with citizen stakeholders. In most, if not in all, of the plans available, the focus is mainly on the outcomes. In fact, at the heart of a conflict management plan should be the user because conflict management plans should be operational manuals – rather than a review of the literature – that help the user to make timely decisions about the use of potential tools under various conditions (related to staff, organisational structure, equipment and gear, etc.). Moreover, as is the case in some other sectors, such as in aviation (FAA 2000), in conflict management, the greatest variability is also the human factor. Similar to the aviation sector (FAA 2000), the factors (such as stress, peer pressure, adaptive skills, physical and motivational fatigue) that affect the performance of people managing conflicts are topics that are yet to be studied by relevant academic disciplines. Speaking about the human factor, leadership has a crucial role in conflict management. As Peter Drucker stated, “only three things happen naturally in organisations: friction, confusion, and underperformance. Everything else requires leadership” (in Sandermoen 2019). In challenging environments, what is most needed in conflict management is leadership. In countries where wildlife agencies have limited capacity to work with local communities or lack the will to do so, this leads to public frustration (Baruch-Mordo et al. 2011). Currently, managing human–bear conflicts is arguably one of the most challenging topics for wildlife managers in the USA (Spencer et al. 2007). If anything, the challenge is even bigger for those involved in managing conflicts in challenging environments, and there are two reasons for that: human–bear conflict is a wicked problem, and applicability of the 17 conflict management tools is limited in resource-limited countries.

Human–Bear Conflict is a Wicked Problem Human–bear conflict is more of a wicked problem than a complex problem in challenging environments. Complex problems are those for which problem solvers agree on what the problem is, but for which there is no consensus on the solutions. On the other hand, for wicked problems, there is disagreement about the problem as well as the solution; planning problems are wicked problems (Rittel & Webber 1973). For wicked problems, the issue is not related to uncertainties in science or failure to consider scientific evidence, but the deep disagreement on values and world views (Balint et al. 2011). Therefore, traditional top-down approaches to conflict management where an agency produces a management plan by getting input only from scientific elites do not provide effective solutions to conflicts. As a result, the problem escalates. Identifying

human–bear conflicts as wicked problems makes it clear that we need to work on the problem definition first to have a common ground regarding conflict management.

The Conflict Management Toolbox Needs Innovation The lack of effective governance arrangements, financial mechanisms, and institutional and technical capacity make human– bear conflict management practically impossible in some resource-limited countries. Only a fraction of the conflict management tools used in North America or Europe is available to the practitioner in resource-limited environments (Table 16.1). This is sometimes the reason for the seeming unwillingness of wildlife authorities in dealing with conflicts in resource-limited countries. Therefore, it is incumbent on the wider conservation community to develop innovative, cost-effective, thoughtful solutions (Can et al. 2014). However, nothing is easy, it seems. Although there is an increasing emphasis on the use of technology in conservation, it will take more than technology to save the world (for more information see Can & Macdonald 2017). Moreover, transferring approaches that work in resource-rich environments (WEIRD countries) to challenging environments rarely works in the long term. Gippoliti et al. (2018) thoughtfully discussed why the Western European approach of large mammal conservation is not exportable to other parts of the world and, in fact, may have negative consequences for wildlife. This applies to conflict management as well because conservation polices – as economic and foreign policies – that are mostly based on utilitarian considerations have a high chance of failure outside WEIRD countries (Can & Macdonald 2018). So is there hope for new ideas, concepts, and approaches? Deeper and contextual evidence will probably come from specific case studies done in resource-limited regions and by going beyond biological insights.

Conflict Management Plans as Logic Models Each plan for human–bear conflict management is a proposed solution to the problem of conflict at a given site, region, or country. One way to produce better conflict management plans is to approach the plan as a logic model (for more information, see Knowlton & Philips 2009). The name may sound fancy, but a logic model is simply a detailed picture of how conflict management will be practiced on the ground. It is a tool that can be used not only for design and implementation but also for evaluation and reporting (W. K. Kellogg Foundation 2004). A logic model presents all available resources, actions that will be carried out with the available resources to solve the problem (i.e. activities), what the activities will deliver (i.e. outputs), the change that will occur as a result of the outputs (i.e. outcomes), and finally what these changes will eventually lead to (i.e. impact) (W. K. Kellogg Foundation 2004; UNDP 2009; Busjeet 2013; Figure 16.3). For every entry in the plan, from activities to impact, the ownership, location, and time are specified; therefore, who

231

Human–Bear Coexistence Table 16.1 List of 17 tools for human–bear conflict management and subjective assessment of their suitability for a resource-rich environment vs. a resourcelimited environment.

Name of the tool/approach

Potential for application in a: Resource-rich environment

Resource-limited environment

Aversive behavioral conditioning

Medium to high

Low to medium

Avoiding negative encounters

High

Low to medium

Awareness raising

High

Low to high

Bear population management

Medium to high

Low

Compensation & insurance programs

Medium to high

Low

Diversionary & supplementary feeding

Medium to high

Low

High

Low to medium

Early warning approaches

Medium to high

Low

Habitat management

Medium to high

Low

Landscape planning

Medium to high

Low

Legislation development & enforcement

Medium to high

Low

Management of attractants

High

Low to medium

Relocation & translocation

Medium to high

Low

Removal of conflict individuals

Medium to high

Low to medium

Traditional methods

Low

Medium to high

Medium to high

Low to medium

Medium to high

Low to medium

*

Education

*

Use of bear deterrents **

Use of physical barriers

* Although education and awareness are often mentioned together in the literature, they are actually two different endeavors. ** Elevated platforms, electric fences, etc.

Human–Bear Conflict Management

A

Resources

Activities

(What will be needed?)

(What will be accomplished?)

Outputs (What will the activities deliver?)

Planned work

Outcomes (What will change in the short/medium/ long term?)

Impact (What will the outcomes lead to?)

B

Intended results Time

Initial state of human–bear conflict

Desired state of human–bear conflict

Figure 16.3 Schematic representation of human–bear conflict management as a logic model. The resources that are available to the practitioners dictate the type of activities that can be reasonably planned to deal with conflicts. Then, outputs are achieved when those activities are accomplished. Outcomes are achieved as a result of the outputs. Finally, outcomes determine the overall impact of the conflict management plan.

does what, where, and when is clearly documented. Outcomes should be specific, measurable, action-oriented, realistic, and timed (known as the SMART acronym). For an NGO, this could be signing a memorandum of understanding with the national wildlife authority, or for a wildlife agency, a change in the number of complaints received from the local community

232

can be examples of short-term outcomes. Impact will be the ultimate state of the world desired by the people who prepared the conflict plan. Thinking of conflict management plans as logic models helps everyone (as many key contributing interest groups as necessary) involved in conflict management planning to visualize and understand how each step contributes to

Principles of Human–Bear Conflict Management

the ultimate intended goal of the management activities: the coexistence of people and bears. Human–bear conflict plans prepared using logic model thinking not only guide the actions on the ground, but also enable better monitoring and evaluation, and indicate where improvements are needed within the plan. Another way to improve conflict management planning is to consider a Prevent–Respond–Recover (P2R) framework and organize the conflict management plan from a threedimensional time perspective: before the conflict events, during the conflict events, and after the conflict events. A conflict event is when there is potential for loss of human life and/or damage to property. The Prevent phase is about how to prevent or reduce conflict before the actual conflict event occurs by using appropriate tools from the conflict management toolbox. This is when it is best to meet local people to build relationships with them, and collect data about the local community, the bears, and the habitat to fill gaps in the existing knowledge. This is also the best time to run awareness-raising and educational initiatives, to become familiar with the relevant legislation about bears, the habitat, and the legal basis of potential interventions that might be necessary in the short and medium time frame. Prioritizing the communities (i.e. specific areas on the map) by considering their recovery potential (socioeconomic conditions) from potential conflict events at this stage will also save time when large-scale conflicts occur. Similarly, the Respond phase relates to how to best prioritize actions and respond to human–bear conflict situations while they are happening. To effectively respond to conflicts, it is vital to understand the impact of the conflict not only on the affected individuals but also on the community as a whole. Making commitments to the local community that will not be followed later on will harm the Recovery phase of conflict management. The ultimate purpose of responding to conflicts is to implement necessary measures to reduce or eliminate the risks caused by bears – in terms of human safety and damage to property – so that conflict does not escalate leading to retaliatory killings of bears. Lastly, the Recovery phase refers to the period when tools such as compensation and insurance programs can be used. The focus of this phase is managing people’s expectations as much as possible to recover from the trauma.

Key Factors for Managing Human–Bear Conflicts in Challenging Environments The following factors play a key role in managing human–bear conflicts in challenging environments. 1. Recognize that conflict is real and is a biological phenomenon happening in the human brain. Individuals, communities, and societies think in very diverse ways about animals; the human–animal relationship cannot be simply explained by solely measuring attitudes and behavior (Pooley et al. 2017). Sometimes social factors

(such as peer group norms) but not the perceived threats from large carnivores such as bears are the reason for killing predator species (Treves & Bruskotter 2014). For the human brain, some problems (such as which bus to take to the city center) activate brain regions involved in logical problem-solving (Shenhav & Greene 2014; Eagleman 2016). Other problems, such as hearing about a bear attack on a neighbor, activate other brain networks, such as those involved in emotion. Many human conflicts are disputes over values but not evidence (Ginges et al. 2007). Research shows that people may not reason instrumentally when values are at stake (Tetlock 2003). This is true for human–human conflicts that are embedded in human–wildlife conflicts. Conflict is a real biological phenomenon that happens in the human brain and neuroscience has recently started to understand the neurological processes behind conflicts (Can & Macdonald 2018). Human–carnivore conflict is rarely, if ever, solely due to the damage received by large predators. When large carnivores are considered, activity in the brain networks that are involved in fear and other emotional processes increases (Can & Macdonald 2018). In brief, conflicts are about how the human brain is wired. Do not get into a fight with it, live with it. 2. Recognize the fact that every human–bear conflict crisis contains the opportunity to make things worse. The first principle of conflict management interventions can be perhaps thought of as the consequentialist maxim “first, do no harm” as is often stated in medicine (Marsh 2014). The planned activities for managing conflicts should not make things worse in the affected community (Chandran 2015). For example, in a community affected by damage caused by bears to apiaries, if the support (monetary or not) that the authorities provide to mitigate the negative effects of damage is not delivered to the right families (for example, due to local power dynamics), this would result in a decline in the public’s trust of the authorities. This, in turn, will make effective conflict management an even bigger challenge in that particular resource-limited environment by making things worse. 3. Recognize the negative effects of not acting when needed or acting too little or too late in significant conflicts. Responding to conflict emergencies, e.g. where there is human fatality due to a bear attack, needs a different set of resources than slow-onset conflict situations. In conflict emergencies, there is a need for a fast flow of support to the affected family and local communities, within days. Planning conflict management across a country and ensuring multi-year resources devoted to conflict management is a dream in resource-limited countries. Nevertheless, management authorities should have short(for example, constructing physical barriers such as elevated platforms), medium-, and long-term goals (such as habitat restoration for bears) in mind.

233

Human–Bear Coexistence

4. Use the capacity to first deal with the most serious type of all conflicts: human death caused by bears. Fatalities due to bear attacks are “conflict emergency situations” and are more common in resource-limited countries. However, to investigate deaths caused by bear attacks requires the involvement of forensic and genetic laboratories (Can et al. 2014). Except in a few countries in Europe and outside North America, wildlife agencies do not have that capacity or resources. Nevertheless, it is crucial to demonstrate to the affected family and the community that the wildlife authorities are doing all they can to prevent future tragedies. Failing to do so will cause the tolerance level toward bears to drop. Expect that this will sooner or later hinder not only conservation of bears but also other conservation initiatives in that particular area. 5. Support families and communities that are affected by conflicts. In poorer parts of the world, there is an absolute need to integrate human–bear conflict management with poverty alleviation strategies. Bear attacks – when not resulting in fatalities – may leave the victims disabled, leaving them unable to perform necessary activities, which in turn affects the whole family within the socioeconomic framework of rural communities (Ratnayeke et al. 2014). Assess what would be the best thing to do to support people, but be informed about cultural sensitivities and beware of unconscious biases (for more details see Brownstein 2019). Providing financial support might not be always an option due to relevant (or lack of ) legislation. If that is the case, try to innovate with a solution that considers what is possible in terms of the local legislation. Providing a government service might be an option (such as providing free transportation to village children to reach school, providing free meals, providing free health service for the elderly, etc.). Prioritize and focus on areas where the conflict is a threat to the viability of bear populations. 6. Be honest about the limits of effectiveness for each conflict management tool. Involving and gaining the trust of the local community is a crucial step in conflict management. For effective conservation, the public needs to trust the bear managers who are acting to protect public safety (Herrero et al. 2011). Sharing responsibility in conflict management is said to be a key element for success in North America (Treves et al. 2006; Matt 2012), but in resource-limited parts of the world this is harder than gaining the trust of the local communities. There is no single magical tool that can end human–bear conflicts. As societies need a full-time police force to maintain law and order, and policing is a never-ending job, human–bear conflict management will also remain a full-time endeavor as long as bears and people coexist. If possible, practice proactive enforcement, i.e. dispensing notices to warn the public about bears, that will likely be more effective for behavioral change toward creating bear-friendly communities (Baruch-Mordo et al. 2011).

234

7. Education should be tailored to the needs of the community and should have specific targets. Education and awareness are the most mentioned tools in conflict literature (Can et al. 2014). Awareness-raising is not education and education is not just giving systematic information. The purpose of education is to achieve the desired behavioral changes in the target population. However, there is not always a direct link between knowledge and actual change in behavior (Baruch-Mordo et al. 2011). Therefore, the development of more effective educational methods is needed. Those involved in designing and running education initiatives should assess and address the values of the target population, and tailor and implement education initiatives within the framework of adaptive management (Treves et al. 2006; Gore et al. 2006, 2008; Spencer et al. 2007; Baruch-Mordo et al. 2011; Reddy et al. 2017). 8. Pay attention to the welfare of wild bears. Ethical dimensions of conservation science tend to be neglected in comparison to other academic fields such as medicine (Paquet & Darimont 2010). The field of conflict management is no exception. Conservation is concerned with populations and populations are emergent properties of individuals. Although it is not easy, there is a need to quantify the welfare implications of human–bear conflict. Welfare science is interested in the behavioral processes that are also of interest to conservation (Swaisgood 2007). Remember that the public more intuitively understands individual animals than abstract populations (Can et al. 2014). 9. Search, document, and try to innovate traditional ways to alleviate bear damage. Actively search for traditional approaches to conflict management, even if they may not be effective in reducing damage. Nevertheless, document and share them with colleagues, members of the IBA and IUCN Bear Specialist Group, and try to brainstorm ideas to innovate damage prevention techniques using locally available materials. 10. Think big but start small. International donor NGOs that fund the local NGOs working in wildlife conservation are in competition with each other to report success stories to their own donors and the general public (i.e. how their projects made a positive impact in the real world). Therefore, by default, most NGOs have to focus on “lowhanging fruits.” Nevertheless, when like-minded people working in different institutions align, remarkable success can be achieved in challenging environments. One such example is from Turkey: the Brown Bear Research and Conservation Project run by Doğa Derneği, a local NGO, with the support of World Animal Protection (formerly known as WSPA). After four years of preliminary work, a one-year project on human–bear conflict was initiated in 2006. Successful demonstration of the use of electric fences to protect orchards, and innovative bear-proof platforms

Principles of Human–Bear Conflict Management

Figure 16.4(A–D) An example of a traditional wooden beehive condominium placed on a cliff to protect beehives from the reach of bears in Turkey (A and B). Elevated bear-proof beehive platforms (C) designed by the author provided a practical and safer method to protect the beehives; electric fences protect the orchards in Turkey (D). (Photos by Ö. E. Can.)

designed by the author to protect beehives in Artvin, Rize, and Erzurum provinces, attracted nationwide interest (Can et al. 2007, 2014; Figure 16.4). As a result, the duration of the project was extended in 2008, and the author became the scientific advisor to the Ministry of Forestry, the national authority on bear research and conflict management. Within the framework of the project, the use of modern tools and techniques such as camera trapping, live trapping and handling bears, and the use of GPS-GSM collars were demonstrated for the first time in Turkey (Lise 2011). Then, the Ministry of Forestry started restoration of brown bear habitats throughout Turkey. This was possible after a paradigm change in forestry practices, as a result of the discussions held within the framework of the project. Moreover, an activity book for

teachers and students prepared within the framework of the project was approved by the Ministry of Education and is used in schools. After hundreds of articles in the media reaching an estimated 10 million people (out of a country population of 73 million), the state television TRT prepared a documentary featuring the project in 2010. The documentary won several international awards and has been aired regularly since then, by now probably reaching every beekeeping family in Turkey. Later, the project team also assisted a Global Environment Facility (GEF)-funded project run by UNDP Turkey, WWF Turkey, and the Ministry of Forestry in 2011 to replicate the project in other parts of the country. The project approach was replicated by smaller NGOs too. In brief, in about six years, a small team of dedicated people from a local NGO, an

235

Human–Bear Coexistence

international donor, and the Ministry of Forestry succeeded in changing the paradigm about human–bear conflict in a resource-limited environment. Why can’t this be replicated in other challenging environments? 11. Be positive, patient, and persistent (the mantra of conflict management in challenging environments). Practitioners, particularly members of NGOs, increasingly find themselves under pressure to report back the success of the initiatives their donors are supporting even in the first year of projects. However, making real impact by working with local communities is harder to achieve than is reported. In the Brown Bear Research and Conservation Project example from Turkey, the use of bear-proof beehive platforms spread through the villages slowly as expected according to the diffusion of innovation theory (for more information, see Rogers 1995). It took significant time for the locals to become used to the innovative idea, but once the psychological resistance to innovation is broken, the new idea spread quickly and bear-proof beehive platforms flourished throughout the villages. The lifespan of typical NGO projects (one to two years) is too short to see the real impact of the projects. Therefore, it is necessary to negotiate with the donor organizations by explaining the diffusion of innovation theory while staying positive, patient, and persistent. 12. Recognize that ignoring conflicts will have consequences. Working on conflict is different from working around conflict, which is acknowledging conflict as an issue, but not actively seeking to engage in its management. If the authorities that are in charge of bear protection and conflict management do not take an active role to prevent conflict situations from escalating, then they will probably have to deal with the consequences for people and bears.

Conclusion A recent study indicates that humans played a major role in the local extirpation and the general extinction of the

References Artelle, K. A., Reynolds, J. D., Treves, A., et al. (2018). Hallmarks of science missing from North American wildlife management. Science Advances 4(3): eaao0167. Balint, P. J., Stewart, R. E., Desai, A. & Walters, L. C. (2011). Wicked environmental problems: Managing uncertainty and conflict. Washington, DC: Island Press. Baruch-Mordo, S., Breck, S. W., Wilson, K. R. & Broderick, J. (2011). the carrot or the stick? Evaluation of education and enforcement as management tools for

236

European cave bear (Ursus spelaeus) (Gretzinger et al. 2019). However, Ötzi the Iceman and his contemporaries coexisted with brown bears some 5300 years ago. Today, humans and bears continue to coexist. However, this coexistence largely depends on changes in human behavior; therefore, like conservation, conflict management is a social and political endeavor rather than just a technical one (Peyton 1994). A wildlife manager is expected to “fix” a conflict situation in very little time. On the other hand, being a researcher provides the luxury of having the time, a relatively independent position to think about conflict, and perhaps a learned ability to think in broader terms. Therefore, it is perhaps more of a moral imperative to state this for this author: We don’t know what the population status of the eight species of bears will be and the level of conflict that will be affecting bears in 50 or 100 years’ time. What is more troubling is that we have no means of knowing about how people’s values will change on the four continents where bears exist in that time frame, but let’s hope that it will be toward a positive direction. An African proverb tells us, “Until the lion learns how to write, every story will glorify the hunter.” Let us hope that future generations, when they look back, won’t think that today’s researchers and practitioners made this mistake.

Acknowledgments I would like to dedicate this chapter to my dad, Professor Vedat Can, who passed away during the preparation of this chapter (I miss you a lot, Dad). I am grateful for Professor Wilfried Buetzler (from Germany) for teaching me how to track bears 22 years ago; Dr. George Schaller (from the USA) and Victor Watkins (from the UK) for persuading me to work on human–bear conflicts; members of the International Association for Bear Research and Management (IBA) and the IUCN Bear Specialist Group for all the discussions we had over the last 20 years. I thank all researchers, game wardens, wildlife managers, decision makers, and local communities that I have worked with around the world for all the conversations we have had.

human–wildlife conflicts. PLoS ONE 6(1): e15681. Boroditsky, L. (2010, July 24). Lost in translation. The Wall Street Journal. Boroditsky, L. (2011). How language shapes thought. Scientific American, 4.

Can, Ö. E. & Macdonald, D. W. (2017). To protect everything, please click here: does a revolution in data collection guarantee one in conservation? Animal Conservation, 20(1): 1–2.

Brownstein, M. (2019). Implicit bias. Available from https://plato.stanford.edu/entries/ implicit-bias/ (accessed March 9, 2019).

Can, Ö. E. & Macdonald, D. W. (2018). Looking under the bonnet of conservation conflicts: can neuroscience help? Biodiversity and Conservation 27(8): 2087–2091.

Busjeet, G. (2013). Planning, monitoring, and evaluation: Methods and tools for poverty and inequality reduction programs. Washington, DC: World Bank. Retrieved from https://openknowledge.worldbank .org/handle/10986/16449

Can, Ö. E., Lise, Y. & Kandemir, İ. (2007). Bees and bears: a review of beekeeperbear conflict in Black Sea Region, Turkey and recommendations for conflict resolution. American Bee Journal, July: 630–636.

Principles of Human–Bear Conflict Management

Can, Ö. E., D’Cruze, N., Garshelis, D. L., Beecham, J. & Macdonald, D. W. (2014). Resolving human–bear conflict: a global survey of countries, experts, and key factors. Conservation Letters 7(6): 501–513.

Hristienko, H. & McDonald, J. E. (2007). Going into the 21st century: a perspective on trends and controversies in the management of the American black bear. Ursus 18(1): 72–88.

Chandran, R. (2015, March 9). It’s broke, so fix it: humanitarian response in crisis. Available from https://cpr.unu.edu/itsbroke-so-fix-it-humanitarian-responsein-crisis.html (accessed April 5, 2015).

Institute of Medicine. (2009). On being a scientist: A guide to responsible conduct in research, 3rd edition. Washington, DC: National Academies Press.

Cohn, J. P. (2002). Environmental conflict resolution. BioScience 52(5): 400–404. Eagleman, D. (2016). The brain: The story of you. Edinburgh: Canongate. Editors. (1957). Approaches to the study of social conflict: introduction by the editors. Conflict Resolution 1(2): 105–110. Elm, E. C. & Diener, M. K. (2007). The language of war in biomedical journals. The Lancet 369: 1.

IUCN SSC BSG. (2019). Principles of human–bear conflict reduction. IUCN SSC Bear Specialist Group. Knowlton, L. W. & Phillips, C. C. (2009). The logic model guidebook: Better strategies for great results. Thousand Oaks, CA: SAGE. Larson, B. M. (2005). The war of the roses: demilitarizing invasion biology. Frontiers in Ecology and the Environment 3(9): 495–500.

Federal Aviation Administration. (2000). FAA system safety handbook. Retrieved from www.faa.gov/regulations_policies/ handbooks_manuals/aviation/risk_ management/ss_handbook/

Lise, Y. (2011). Doğa Konuşmaları Yaban Hayat. Yesil Atlas, 8.

Ginges, J., Atran, S., Medin, D. & Shikaki, K. (2007). Sacred bounds on rational resolution of violent political conflict. Proceedings of the National Academy of Sciences 104(18): 7357–7360.

Matt, C. (2012). 4th International Human– Bear Conflicts Workshop Summary. Available from www.bearbiology.org/wpcontent/uploads/2019/08/4th-humanbear-conflict-workshop-summary.pdf

Gippoliti, S., Brito, D., Cerfolli, F., et al. (2018). Europe as a model for large carnivores conservation: is the glass half empty or half full? Journal for Nature Conservation 41: 73–78.

OED. (2015). Conflict. Oxford English Dictionary. Oxford, UK: Oxford University Press. Retrieved from www .oed.com/view/Entry/38898?result=1& rskey=uilnn5&

Gore, M. L., Knuth, B. A., Curtis, P. D. & Shanahan, J. E. (2006). Education programs for reducing American black bear–human conflict: indicators of success? Ursus, 17(1): 75–80.

O’Sullivan, N. J., Teasdale, M. D., Mattiangeli, V., et al. (2016). A whole mitochondria analysis of the Tyrolean Iceman’s leather provides insights into the animal sources of Copper Age clothing. Scientific Reports 6(1). doi:10.1038/srep31279

Gore, M. L., Knuth, B. A., Scherer, C. W. & Curtis, P. D. (2008). Evaluating a conservation investment designed to reduce human–wildlife conflict: evaluating conservation investment. Conservation Letters 1(3): 136–145. Gretzinger, J., Molak, M., Reiter, E., et al. (2019). Large-scale mitogenomic analysis of the phylogeography of the Late Pleistocene cave bear. Scientific Reports 9(1). doi:10.1038/s41598-019-47073-z Hardin, G. (1968). The tragedy of the commons. Science 162(3859): 1243–1248. Herrero, S., Higgins, A., Cardoza, J. E., Hajduk, L. I. & Smith, T. S. (2011). Fatal attacks by American black bear on people: 1900–2009. The Journal of Wildlife Management 75(3): 596–603.

Marsh, H. (2014). Do no harm: Tales of life, death and brain surgery. London: Weidenfeld & Nicolson.

Paquet, P. C. & Darimont, C. T. (2010). Wildlife conservation and animal welfare: two sides of the same coin. Animal Welfare 19(2): 177–190. Paulus, M. P., Potterat, E. G., Taylor, M. K., et al. (2009). A neuroscience approach to optimizing brain resources for human performance in extreme environments. Neuroscience & Biobehavioral Reviews 33(7): 1080–1088. Peterson, M. N., Birckhead, J. L., Leong, K., Peterson, M. J. & Peterson, T. R. (2010). Rearticulating the myth of human–wildlife conflict. Conservation Letters 3(2): 74–82. Peterson, M. N., Peterson, M. J., Peterson, T. R. & Leong, K. (2013). Why

transforming biodiversity conservation conflict is essential and how to begin. Pacific Conservation Biology 19(2): 94. Peyton, B. (1994). Conservation in the developing world: ideas on how to proceed. Bears: Their Biology and Management 9: 115. Pooley, S., Barua, M., Beinart, W., et al. (2017). An interdisciplinary review of current and future approaches to improving human–predator relations: improving human–predator relations. Conservation Biology 31(3): 513–523. Presnall, C. C. (1943). Wildlife conservation as affected by American Indian and Caucasian concepts. Journal of Mammalogy 24(4): 458–464. Ramsbotham, O., Woodhouse, T. & Miall, H. (2011). Contemporary conflict resolution, 3rd edition. Bristol: Polity. Ratnayeke, S., Van Manen, F. T., Pieris, R. & Pragash, V. S. J. (2014). Challenges of large carnivore conservation: sloth bear attacks in Sri Lanka. Human Ecology 42: 467–479. doi:10.1007/s10745-014-9643-y Reddy, S. M. W., Montambault, J., Masuda, Y. J., et al. (2017). Advancing conservation by understanding and influencing human behavior: human behavior and nature. Conservation Letters 10(2): 248–256. Redpath, S. M., Young, J., Evely, A., et al. (2013). Understanding and managing conservation conflicts. Trends in Ecology & Evolution 28(2): 100–109. Redpath, S. M., Bhatia, S. & Young, J. (2015). Tilting at wildlife: reconsidering human– wildlife conflict. Oryx 49(2): 222–225. Rittel, H. W. J. & Webber, M. M. (1973). Dilemmas in a general theory of planning. Policy Sciences 4(2): 155–169. Rogers, E. M. (1995). Diffusion of innovations, 4th edition. New York, NY: Free Press. Sandermoen, K. (2019). Management consulting (is not) for dummies. Alosen, Switzerland: Sandermoen Publishing. Shenhav, A. & Greene, J. D. (2014). Integrative moral judgment: dissociating the roles of the amygdala and ventromedial prefrontal cortex. Journal of Neuroscience 34(13): 4741–4749. Spencer, R. D., Beausoleil, R. A. & Martorello, D. A. (2007). How agencies respond to human–black bear conflicts: a survey of wildlife agencies in North America. Ursus 18(2): 217–229.

237

Human–Bear Coexistence

Swaisgood, R. R. (2007). Current status and future directions of applied behavioral research for animal welfare and conservation. Applied Animal Behaviour Science 102(3–4): 139–162.

238

human–wildlife conflicts: a review. Human Dimensions of Wildlife 11(6): 383–396.

Tetlock, P. (2003). Thinking the unthinkable: sacred values and taboo cognitions. Trends in Cognitive Sciences 7: 320–324.

UNDP. (2009). Handbook on planning, monitoring and evaluating for development results. New York, NY: United Nations Development Programme.

Treves, A. & Bruskotter, J. (2014). Tolerance for predatory wildlife. Science 344(6183): 476–477.

United Nations. (2004). World population to 2300 (No. ST/ESA/SER.A/236). New York, NY: United Nations.

Treves, A., Wallace, R. B., Naughton-Treves, L. & Morales, A. (2006). Co-managing

Wilson, S. M. (2016). A guidebook to human–carnivore conflict: Strategies and

tips for effective communication and collaboration with communities. Ljubljana, Slovenia: Slovenia Forest Service – LIFE DINALP BEAR project. W. K. Kellogg Foundation. (2004). W.K. Kellogg Foundation 2004.pdf. Michigan, USA: W.K. Kellogg Foundation. WSPA. (2009). Principles of human–bear conflict management. Istanbul: World Society for the Protection of Animals. Zedrosser, A. (2019). The major challenges facing bears. International Bear News 28(2): 2.

Chapter

17

Patterns of Bear Attacks on Humans, Factors Triggering Risky Scenarios, and How to Reduce Them Vincenzo Penteriani*, Giulia Bombieri*, Marı´a del Mar Delgado, Thomas Sharp, Koji Yamazaki, Harendra Singh Bargali, Nishith Dharaiya, Ashish Kumar Jangid, Ravi Kumar Sharma, Babu Ram Lamichhane, Shyamala Ratnayeke, Ivan Seryodkin, Himanshu Shekhar Palei, Ashok Subedi, Hu¨seyin Ambarlı, Jose´ Marı´a Fedriani, Pedro Jose´ Garrote, Klemen Jerina, Ilpo Kojola, Miha Krofel, Prakash Mardaraj, Mario Melletti, Andre´s Ordiz, Paolo Pedrini, Eloy Revilla, Luca Francesco Russo, Veronica Sahle´n, Christopher Servheen, Ole-Gunnar Støen, Jon E. Swenson, and Tom Smith

Introduction The media and scientific literature are increasingly reporting an escalation of large carnivore attacks on humans, mainly in the so-called developed countries, such as Europe and North America (Penteriani et al. 2016; Bombieri et al. 2018a). Although large carnivore populations have generally increased in developed countries (Chapron et al. 2014), increased numbers are not solely responsible for the observed rise in the number of attacks. For example, recent research has shown that people frequently engage in risk-enhancing behaviors that can increase the probability of a risky encounter and a potential attack, and perhaps even alter carnivore behavior (Penteriani et al. 2016, 2017; Garrote et al. 2017). Of the eight bear species inhabiting the world, two (i.e. the Andean bear and the giant panda) have never, or very rarely, been reported to attack humans, whereas the other six species have: sun bears Helarctos malayanus, sloth bears Melursus ursinus, Asiatic black bears Ursus thibetanus, American black bears Ursus americanus, brown bears Ursus arctos, and polar bears Ursus maritimus. These species occur across four continents (Asia, Europe, North and South America) characterized by a huge range of social and cultural practices, e.g. from increasing leisure activities in bear areas in developed countries to daily forest works in developing countries. Such differences in the use of bear habitats by people may determine that different scenarios trigger bear attacks on humans around the world. However, even if the motivations that determine human presence in bear countries and risky encounters with bears are diverse, some triggering factors might be common in activating bears’ dangerous reactions toward people, e.g. inappropriate human behaviors when sharing the landscape with bears or when encountering them at close range. This chapter provides insights into the causes, and as a result the prevention, of bear attacks on people. Prevention and information that can encourage appropriate human *

These authors contributed equally to this work.

behavior when sharing the landscape with bears are of paramount importance to reduce both potentially fatal human– bear encounters and their consequences to bear conservation.

Methods We reviewed scientific/gray literature and analyzed personal databases on bear attacks on humans available from 1980 to 2018. Moreover, we also searched for PhD/MSc theses and webpages on bear attacks on humans. In addition, we collected news reports from online newspapers to complete the data set obtained from the above-mentioned sources. To this aim, for each bear species and area, we searched for news articles on Google on an annual basis using the following combination of words: “species name” + “attack” and “species name” + “attack” + “human.” To prevent duplicate records in the data, we cross-checked information such as date, locality, and human characteristics. When possible, for each attack, we recorded the following information: (1) period of the attack, i.e. year, month, and time of the day; (2) location; (3) outcome, i.e. human injury or death; (4) characteristics of the person/ party and bear involved in the attack, e.g. age and sex of both the person and the bear; (5) human activity at the time of the attack; and (6) the attack scenario, i.e. the factor(s) that could have triggered the attack.

An Overview of Bear Attacks in the World Sun Bear The sun bear, the smallest ursid in the world, is found in South-East Asia, where few (n = 11) attacks on humans have been reported. Although the available information is scarce and incomplete (Sethy & Chauhan 2013), attacks seem to be extremely rare and mainly the consequence of sudden encounters (i.e. with the bear being inadvertently surprised at a close distance). Such encounters mainly occur when people venture inside the forest for different purposes (Sethy & Chauhan 2016). For example, in Indonesia most attacks happen to

239

Human–Bear Coexistence

Figure 17.1 The sloth bear’s aggressively defensive nature may be the result of having coevolved with large predators, such as tigers and leopards, which are known to occasionally prey upon this bear. Actually, sloth bears are strong animals that possess large canines and claws, which make them difficult prey. (Photo by Ayan Sadhu.) (A black and white version of this figure will appear in some formats. For the color version, please refer to the plate section.)

people working in the forest on a daily basis, such as rubber harvesters (55%, n = 6), whereas fewer cases occurred to people who work in crop fields and collect non-timber forest products (Windler 2014). In general, encounters with these bears are mainly non-fatal.

Sloth Bear Sloth bears are known for their aggressive behavior toward humans (Burton 1856; Anderson 1957). Although the total number of people seriously injured or killed by sloth bears during a given year is not known, within one Indian state (Madhya Pradesh, central India), 48 sloth bear-inflicted fatalities and 687 maulings were documented between 1989 and 1994 (Rajpurohit & Krausman 2000). This accounts for an average of six deaths and 115 maulings per year in this Indian state. Because sloth bears occupy 19 Indian states, as well as Nepal, Sri Lanka, and possibly Bhutan, this species might be responsible for more attacks on humans than all other seven species of bears combined. The sloth bear’s motivation to attack seems to be mainly defensive. There has never been a documented predatorial attack (Sharp et al. 2017), although there have been several cases of sloth bears consuming portions of their victims (Bargali et al. 2005; Akhtar 2006). This appears to be the result of opportunistic behavior after the attack, rather than the motivation for the attack. The sloth bear’s aggressively defensive nature may be the result of having coevolved with large predators, namely tigers Panthera tigris and leopards Panthera pardus, which are known to occasionally prey upon sloth bears (Littledale 1889; Fenton 1909; Kurt & Jayasuriya 1968; Laurie & Seidensticker 1977). Sloth bears are not adept climbers and are often in scrub jungle or grasslands, where climbing is not an option. They are not particularly fast runners, either.

240

However, they are exceptionally strong animals that possess large canines and claws. Although no match for a tiger in an extended encounter, sloth bears can make themselves a particularly difficult prey (Figure 17.1). In addition to their physical attributes, sloth bears make use of intimidation tactics, such as charges coupled with vocalizations and making themselves appear larger with a bipedal stance (Bargali et al. 2005; Ratnayeke et al. 2014; Sharp et al. 2017). This aggressive– defensive nature may serve to deter predators, but impedes conservation efforts whenever they attack humans (Akhtar 2006; Dharaiya et al. 2016). To better understand this type of human–sloth bear conflict, we compiled information on 1169 attacks that took place in the three countries with extant sloth bear populations: India, Nepal, and Sri Lanka. We also considered the reported results from individual studies conducted throughout the sloth bear’s distribution. We found that the spatiotemporal patterns of these attacks varied across studies. Whereas some studies reported a spike in attacks during the monsoon season, others reported more attacks during the dry season (Rajpurohit & Krausman 2000; Bargali et al. 2005; Ratnayeke et al. 2014; Garcia et al. 2016; Dhamorikar et al. 2017; Sharp et al. 2017). These differences may be related to seasonal and daily patterns of human use of forests, or to proximity between bears and people in highly fragmented bear populations. For example, in Bilaspur (India), attacks mainly occur in villages and agricultural fields (Bargali et al. 2005), whereas in other study areas, the majority of attacks occur in forests (e.g. Ratnayeke et al. 2014; Dhamorikar et al. 2017). The high number of attacks during monsoon may be because people start agricultural practices in crop fields, collect mushrooms from the forest, and use forests for livestock grazing at the onset of this season. In monsoon, there is increased vegetation cover as well. Such a combination of different factors and the increased disturbance

Patterns of Bear Attacks

in and around sloth bear habitats considerably augment the probability of risky encounters with a sloth bear. The time of day when attacks occurred also varied a great deal. Although some studies suggested that dawn and dusk were the most dangerous times to be moving through sloth bear country (Bargali et al. 2005; Mardaraj & Dutta 2011; Mardaraj 2015; Debata et al. 2017; Dhamorikar et al. 2017; Sharp et al. 2017), others reported that most attacks occurred during the middle of the day (Ratnayeke et al. 2014; Garcia et al. 2016). There is little doubt that local factors, including human activity patterns, play a large role in the spatiotemporal patterns of sloth bear attacks. For example, although sloth bears are generally known to be more active at dusk and during the night, in many places several attacks occurred during the day. This may be partly due to resting bears being disturbed, or surprised, rather than the presence of active bears; actually, very few attacks happen during daylight hours in areas that are known to have many natural caves and hollows. A large portion of attacks (42%) took place when humans were active in the forest (e.g. collecting forest products, walking, etc.). For example, in central India, people collect tendu leaves (Diospyros melanoxylon) and mahua flowers and seeds (Madhuca indica) for commercial use, as well as sal leaves (Shorea robusta), bamboo (Dendrocalamus strictus), chironji (Buchanania lanzan), and wild mushrooms, which increases the probability of sudden encounters between people and sloth bears. The second highest category (25%) involved persons who were farming or caring for orchards. These types of activities are more often performed by men than by women or children, which likely explains why adult males were found to be involved in the majority of attacks (87%). Perhaps surprisingly, the third highest number of attacks (15%) took place when humans were defecating in the forest. This tends to occur in areas which are often frequented for this purpose. The reason for this type of attack is unclear. It is possible that fecal odours attract sloth bears to areas often used for defecation. However, this is purely speculative at this point, although it is interesting to note that the smell of human feces is known to also attract both brown and American black bears (T. Smith, unpublished data). Sudden encounters triggering defensive–aggressive attacks accounted for 47% of the attacks. Often the victim was unaware of the sloth bear’s presence until they saw the bear charging from just meters away (Figure 17.2). Like most mammals, sloth bears are protective of their young and, of the 1169 attacks compiled for this chapter, 22% involved females with cubs, although several studies reported this number to be higher (Ratnayeke et al. 2014; Sharp et al. 2017). Most studies on sloth bear attacks report that human group size plays a role in the likelihood of attack (Ratnayeke et al. 2014; Sharp et al. 2017). We found that nearly half of all attacks involved a single person (46%). However, reported group size can be misleading, because if people are spread out, the bear may perceive each individual as a single person, rather than as

Figure 17.2 The main scenario of sloth bear attacks on humans is sudden encounters in dense forests, where the victim is often unaware of the bear’s presence. (Photo by Luxshmanan Nadaraja.)

part of a group. Ratnayeke et al. (2014) found that a human companion 13 years old, 94%), which is likely explained by the same phenomena as the sex bias. Most of the attacks recorded were the result of sudden encounters (86%), with bears reacting aggressively when surprised at close distances. Notably, we found that at least 30% of such encounters occurred while bears were feeding on crops or other products in people’s fields or orchards. Such a scenario has been found to be the main attack circumstance in several studies, with people encountering bears at close range and consequently being attacked when visiting or working in their crop fields or orchards (Nabi et al. 2009a; Tak et al. 2009; Rasool et al. 2010; Charoo et al. 2011). Other reported scenarios involve people entering dense forest to collect wood or other forest products, or to graze livestock (Tak et al. 2009; Lal Moten et al. 2017; Ali et al. 2018). The number of bear attacks has been increasing recently in Nepal (Acharya et al. 2016), India (Kashmir; Nabi et al. 2009b), and Japan (Yamazaki 2010). This is mostly due to human population expansion, deforestation, and destruction and fragmentation of bear habitat in most Asian countries (Japan Bear Network 2006). In contrast, the reason for increasing bear attacks in Japan is quite different, because the population and range of bears have been expanding in recent decades, due to habitat recovery as a result of aging human population and depopulation (Yamazaki 2004; Yamazaki & Sato 2014). From statistics supplied by the Ministry of Environment in Japan, 851 people were attacked (including 13 deaths) by bears in the last 10 years (2008–2017). Although the frequency of attacks by Asiatic black bears is relatively high, these attacks rarely result in fatalities (8%). Similar or lower fatality rates are reported in Nepal (Acharya et al. 2016), India–Kashmir (Nabi et al. 2009b; Rasool et al. 2010; Shah et al. 2010), and Japan (Akiyama et al. 2017). Although fatalities are rare, injuries are frequently serious, as bears often stand upright and first attack the person’s neck and face using their claws, causing damage

242

such as bone fractures and deep tissue lacerations (Nabi et al. 2009b; Rasool et al. 2010; Oshima et al. 2018). As a consequence, many victims must undergo multiple and complex facial reconstructive surgeries (Rasool et al. 2010; Shah et al. 2010). In the northern part of Honshu Island, Akita Prefecture, Japan (Yamazaki 2017; Oshima et al. 2018), four local residents who were gathering bamboo shoots were killed and partially eaten by bears in 2016, and another local resident was attacked and partially eaten by a bear(s) in 2018. As a result, local attitudes toward bears have drastically changed within this prefecture and have become more negative, resulting in a total of 1676 bears being killed as nuisance animals between 2016 and 2018. These attacks have had a bad influence even in other prefectures, and bear management is facing serious difficulties in many areas.

American Black Bear Outnumbering the other two species of North American bears 10-fold, the American black bear is the most ubiquitous ursid on the continent (Herrero 2018). Not surprisingly, black bears account for the majority of human–bear incidents in North America (Penteriani et al. 2016). Indeed, our search resulted in 373 attacks reported between 1980 and 2016, and only 11% were fatal. In a recent study of human–bear conflict in Alaska, Smith and Herrero (2018) reported that black bears, which outnumber brown bears three to one in the state, were responsible for only 14% (89 of 638) of bear attacks reported from 1880 to 2016. This suggests a tolerance for humans not shared by brown bears. Herrero (1972) suggested that this reluctance to engage with people is the result of their unique evolutionary past. Black bears evolved in the densely forested regions of North America and could often resolve conflict by either climbing a tree or disappearing into the dense understory (Herrero 1972). Therefore, when suddenly confronted by people, they frequently flee. Indeed, Smith and Herrero (2018) found in Alaska that brown bears accounted for nearly six times more attacks than black bears (508 vs. 89), resulting in 47 brown bear-inflicted fatalities compared to just five due to black bears (a 9.4-times higher rate for brown bears). Nonetheless, on rare occasions, when black bears attacked and killed people, 90% of them were deemed predatory (Herrero et al. 2011). From 1900 to 2009, Herrero et al. (2011) documented 63 fatal black bear attacks in the United States and Canada. They also found that black bear-inflicted fatalities were highly correlated with human population growth in those countries, suggesting that the more people enter black bear habitat, the more likely it is that a conflict will arise. Although the vast majority of black bears clearly avoid conflict with humans, a few bears apparently perceive people as prey and attempt to take them. Herrero and Higgins (2003) identified predatory black bears by a series of behaviors: searching, following and testing,

Patterns of Bear Attacks

attacking (capturing), killing, sometimes dragging a person, sometimes burying, and often feeding upon a person. By analyzing patterns of predatory attacks by all large carnivore species in North America, Penteriani et al. (2017) found that the black bear was the third species most frequently involved in this kind of encounter (16%), after cougars and coyotes, and, as expected, the targets of predatory attacks were the most vulnerable individuals, namely children and lone people. Black bears are powerful predators quite capable of inflicting severe injury and death on humans. Herrero et al. (2011) speculated that black bears likely do not prey on people more often than they do because bears with those genes have been consistently culled from the gene pool. Importantly, fatal attacks by black bears are fundamentally different than those of brown bears because browns are rarely predatory, whereas black bear-inflicted fatalities are almost always the result of predation. Unlike brown bears, black bears have been often reported to cause conflicts and, more rarely, injure people in urban areas across North America (Bombieri et al. 2018b). In such environments, risky encounters with this species are mainly due to the presence of dogs or anthropogenic food (e.g. a bear is surprised by a person while feeding on pet food or trash in the yard and reacts aggressively). Probably due to the forest-obligate nature of the species, attacks in urban areas are more likely to occur where the vegetation cover is more abundant and less fragmented and far from buildings and roads. Half of such encounters have been found to occur at night, especially in areas where the artificial illumination is scarce (Bombieri et al. 2018b).

Brown Bear There are always conflicts wherever humans and brown bears commingle. Several studies have shown that increases in the number of brown bear conflicts are roughly proportional to human population growth in North America (Herrero et al. 2011; Sharp et al. 2017; Smith & Herrero 2018). Although long absent from many localities, people still harbor an inordinate fear of brown bears; a fear that is deeply rooted in our past history, as explained by Wilson (1984): The brain evolved into its present form over a period of about 2 million years, from the time of Homo habilis to the late stone age of Homo sapiens, during which people existed in hunter-gatherer bands in intimate contact with the natural environment. The smell of water, the hum of a bee, the directional bend of the plant stalk mattered. The naturalist’s trance was adaptive: the glimpse of one small animal hidden in the grass could make the difference between eating and going hungry in the evening. And a sweet sense of horror, the shivery fascination with monsters and creeping forms that so delights us today even in the sterile hearts of our cities, could see you through to the next morning. Organisms are the natural stuff of metaphor and ritual. Although the evidence is far from all in, the brain appears to have kept its old capacities, its channelled quickness. We stay alert and alive in the vanished forests of the world.

This innate fear is kept alive by the media, which often focuses an inordinate amount of attention on attacks by brown bears, thus fueling fear and undermining conservation efforts (Penteriani et al. 2016; Bombieri et al. 2018a). Human–brown bear conflicts occur throughout their range, and though few in number, have profound impacts on both the people and bears involved (Conover 2008; Penteriani et al. 2016). A study by Bombieri et al. (2019) investigated circumstances of brown bear attacks that occurred worldwide between 2000 and 2015 and highlighted common characteristics, as well as differences in attack patterns between regions of the world with different sociocultural backgrounds and histories of coexistence with this species. According to this study, people involved in attacks are almost exclusively adults (99%) and human fatalities are rare (15%). Globally, attacks are mainly the result of an encounter with a female bear with cubs (47%; Figure 17.3A), followed by sudden encounters with other bear classes (20%), dog presence (17%; Figure 17.3B), bear attacking after being shot or trapped (10%), and predatory attacks (5%). At the moment of the attack, half of the people were engaged in leisure activities (e.g. hiking, picking berries or mushrooms, camping, fishing), whereas 28% of the attacked people were working outside, i.e. farming, guarding livestock, or logging, or doing wildlife-related fieldwork, and 22% were hunting. Although patterns are similar in many regions across the wide distribution range of brown bears, local differences in attack scenarios also exist. For example, in Romania, where traditional semi-subsistence agriculture and livestock husbandry are common, half of the people involved in attacks were shepherds, cattle herders, and farmers (Bombieri et al. 2019), whereas most (75%) of the attacks that occurred in Scandinavia between 1977 and 2016 involved hunters (Støen et al. 2018). Hunters are often involved in bear attacks, because they are trying not to make noise while hunting and may consequently surprise bears, thus triggering an attack. Other factors connected to hunting may also explain the higher risk of an attack. In Scandinavia most (73%) of the hunter casualties had shot at the bear at short range (average 8 ± 11 m) before being injured and hunting dogs harassed the bears in 77% of the incidents involving hunters (Støen et al. 2018). A recent study by Smith and Herrero (2018) chronicled 135 years of human–bear conflict involving nearly 700 incidents in Alaska, of which 78% involved brown bears. The findings of this study, although regional, offer useful insights regarding attacks, because brown bears react similarly to human confrontation regardless of locality, even though regional differences in bear responses to humans have been documented. For example, Smith et al. (2008) reported that brown bears in Alaska respond somewhat differently toward humans in coastal areas (areas of abundant resources that support highdensity bear populations, which are more tolerant of humans at very close ranges) compared to brown bears in interior areas (areas relatively scarce in resources supporting low-density populations that are more intolerant of humans). Additionally,

243

Human–Bear Coexistence

(Smith & Herrero 2018; n = 1550), those without a deterrent (i.e. firearms, bear spray, etc.) resorted to desperate and ineffective evasive actions, such as running, tree-climbing, rockthrowing, and attempting to fight the animal bare-handed, strategies that rarely worked. Importantly, having a deterrent did not guarantee that the person would avoid injury; but those with deterrents suffered less injury, and less-severe injuries, than those without. Of particular note, those persons with bear spray largely escaped injury (98%), whereas persons bearing firearms suffered more (24% were injured; Smith et al. 2012). Finally, Smith and Herrero (2018) found that attacking bears focused on the victim’s head and neck regions, a finding which suggests that Herrero (2018) was justified in promoting a defensive posture of lying face-down, hands laced over the back of the neck and legs spread to provide the ability to thwart a bear’s effort to flip the body over and face up. Coming to the aid of an attack victim almost always terminated the attack, perhaps because the bear felt outnumbered and left. Although domestic dogs could be helpful in protecting persons from an attack (47.5%), some abandoned their masters (40%). A few dogs (12.5%) reportedly initiated the attack.

Polar Bear

Figure 17.3 Two of the main brown bear attack scenarios: (A) defensive reaction of a female with cubs; and (B) unleashed dogs harassing bears. (Photos by Ivan Seryodkin.)

it has been posited that heavily hunted brown bear populations become more wary of people, although Swenson (1999) failed to find definitive evidence of that. In general, Smith and Herrero (2018) found that increases in the number of brown bear attacks in Alaska over a 135-year period were strongly correlated with the growth of Alaska’s human population. Similarly, when brown bear attack rates were regressed against a simple index created by combining both brown bear and human densities within a given region of the world, they were found to be positively correlated (data adapted from Bombieri et al. 2019). When bear encounters occurred, bears disproportionately attacked single persons, so grouping together decreased the likelihood of an attack. A similar finding was reported by Penteriani et al. (2016) and Garrote et al. (2017). Group size may yield this effect because larger groups are noisier and alert bears to approaching humans (Herrero et al. 2011). Of the persons involved in bear conflicts in Alaska in this study

244

Human–polar bear conflicts have existed for as long as both species have shared the Arctic. Without a written language, these conflicts went unrecorded by indigenous people, but undoubtedly occurred with some regularity. The earliest written accounts of human–polar bear conflict are sporadic and were recorded by European mariners who kept journals. Among the earliest of these is an account of Dutch mariner William Barents (1594), who attempted to capture a polar bear to bring back alive to Holland (Wikipedia 2019). Unfortunately, the bear broke loose aboard ship, rampaged through the vessel, and was killed in the process. Additionally, Barents wrote a year later that two men in his expedition were attacked and killed by a polar bear, the earliest polar bear-inflicted fatalities we know of on record. More recently, Charlie Brower of Barrow (Alaska) witnessed a polar bear casually crush the skull and ribs of a Native guide in 1883 (Brower 1942). Such graphic accounts have unquestionably fuelled the widely held belief that it is fundamental to a polar bear’s nature to “stalk and kill humans” (Ramachandran 2009). Even though brutal accounts of polar bears ravaging humans dot the pages of history books, one must ask if they truly deserve the name “stalker and killer of humans.” Human–polar bear attack research does not support that claim. In recent years, a number of studies have documented the nature of human–polar bear conflicts (Fleck & Herrero 1988; Clark 2003; Wilder et al. 2017). Both Clark (2003) and Fleck and Herrero (1988) reported that most human–polar bear interactions do not result in injury, and when persons were injured, rates of interactions resulting in injury were

Patterns of Bear Attacks

low (2% and 5%, respectively). Specifically, Clark (2003) reported that only one bear-inflicted injury occurred in 53 polar bear–human interactions. Fleck and Herrero (1988) documented 373 aggressive polar bear–human interactions, of which only 10 bear-inflicted injuries occurred. Because Wilder et al. (2017), reported only on polar bear attacks, their data are not comparable to the previous studies that included non-injurious interactions. However, over a 144-year period, Wilder et al. (2017) documented only 73 attacks on humans by polar bears across their entire range, and in those attacks 20 persons were killed and another 63 injured. By contrast, Smith and Herrero (2018) reported nearly the same number of attacks for brown bears in Alaska in a single decade (2000–2009). The record is clear that polar bears injure and kill far fewer humans that do black and grizzly bears (Fleck & Herrero 1988). Although some may point out that the interaction rates between polar bears and humans are far lower than those of black bears and grizzlies, injuries are low and deaths extremely rare where polar bears and people commingle considerably (e.g. Churchill, Manitoba; Kaktovik, Alaska). In all of these human–polar bear interactions, predation was rare, but when people were killed, predation was considered the primary motivation (Wilder et al. 2017).

The “Take in the Bear Home” Message: How to Avoid and Survive Bear Attacks and Specific Measures of Risk Reduction Sloth Bear Information on sloth bear attacks suggests that being in a group and making noise while moving through sloth bear habitats helps reduce the likelihood of startling a bear at close quarters, giving it the opportunity to leave the area without incident. Bear spray and guns are not available for protection to most who live in sloth bear country and many attacks occur too quickly for weapons to be used effectively. However, villagers carrying a heavy club or walking stick have been able to drive off an attacking bear, and this might prove to be especially effective when more than one person acts in concert (Ratnayeke et al. 2014; Sharp et al. 2017). If a sloth bear is observed, but is not aware of the observer, people should slowly back away, giving the bear as much space as possible. Running has been reported to trigger a chase response, and several runners have been pursued and killed by sloth bears (Sharp & Sonone 2011; Ratnayeke et al. 2014; Sharp et al. 2017). Yelling or throwing stones at a nonaggressive sloth bear may also elicit an attack. Although some have reported that the sloth bear retreated when fought, Sharp et al. (2017) reported that those who fought were more likely to be killed than people that played dead. The safest response to an attack might highly depend on factors such as the bear motivation to attack and the human

characteristics as well as group size. A person that is attacked may decrease the risk of injury or death by using the protective position recommended by Herrero (2018) for grizzly bear attacks: lying face down on the ground with hands locked behind the neck and arms protecting the face. The effectiveness of this defensive position for sloth bear attacks is yet untested; however, most sloth bear attacks are defensive, cause injuries to the face and head, and limited data suggest that the attack does not persist when the bear perceives that the threat has faded. Temporal patterns of sloth bear attacks across different geographic areas suggest that peaks in specific types of human activities increase the potential for human–sloth bear encounters and attacks. These activities may vary by location and managers should work with local people to seek solutions. For example, at locations where bears frequent village compounds and agricultural fields, villagers could be encouraged and financially supported to construct toilets and to use extreme caution moving around when bears are active (Jangid & Sharma 2018). Similarly, avoiding the collection of “nontimber forest products” during twilight hours may reduce the risk of encountering active sloth bears. Moving in groups and advertising one’s presence via sound and loud conversation will give resting bears an opportunity to leave the area. In areas where bears reside within or on the periphery of human settlements, cultivating crops (e.g. maize, ground nuts) that attract bears will increase the likelihood of attacks, let alone crop losses. Training programmes aimed at sustainable livelihoods, including support for alternative forms of agriculture and income to reduce the dependency of local communities on forest resources could help to keep sloth bears and people well-separated. Proper garbage/waste management practices should be promoted in those areas where sloth bears are attracted and approach garbage sites to feed on remnants of fruits and edible materials (A. Jangid, unpublished data). In addition, provisioning of edible materials in remote temples should be stopped, so the encounters can be reduced near temple sites. Indeed, a few cases have been reported of bears breaking in and raiding houses and small temples located inside forests to feed on edible products such as oils left by pilgrims as offerings (A. Jangid, unpublished data; Singh et al. 2017; Jangid & Sharma 2018), alarming the villagers. Finally, for bears that have little alternative but to survive in forest fragments surrounded by agriculture and human settlements, guidelines and policies for safely trapping and relocating them may be the most feasible option (Figure 17.4).

Asiatic Black Bear There are no easy solutions to reduce Asiatic black bear attacks on humans. However, public education about bears (i.e. bear habitat, behavior, and ecology) to local residents and people engaging in various activities in the forest can be an effective approach (Japan Bear Network 2006; Jamtsho &

245

Human–Bear Coexistence

Figure 17.4 Trapping and relocating conflictual sloth bears might represent an alternative to retaliatory killing. (Photo by Ashish Jangid.)

Wangchuk 2016). Because attacks often occur on farmers working on crop fields or orchards, suggesting that bears are attracted by agricultural products, preventive measures aimed at protecting such areas might prove effective to avoid both crop raiding and potentially dangerous encounters between farmers and bears (Jamtsho & Wangchuk 2016). Moreover, people should be cautioned about the possibility of encountering bears when working in their crop fields or moving in the forest and should be provided information on how to avoid being attacked. Moving in a group and making noise while moving in areas with poor visibility might help the bear notice human presence and leave the area to avoid the encounter (Rasool et al. 2010). Given the evident defensive strategy of attacking the head and face region commonly adopted by Asiatic black bears, if a bear reacts aggressively when encountered, people should not fight back, but protect these sensitive regions and adopt a passive position. It has been shown, indeed, that the attack terminated spontaneously in most of the cases, with the bear leaving immediately once the victim was overpowered (Rasool et al. 2010; Shah et al. 2010). Also, increasing the distribution/availability of bear avoidance equipment, such as pepper spray, at affordable prices could also be effective. Once a bear attack has occurred, comprehensive on-site verification and sharing of the obtained information among related organizations is also very important to prevent similar types of incidents in the future.

American Black Bear Smith and Herrero (2018) and Herrero and Higgins (2003) identified a number of insights regarding safety in black bear country. These findings fall into four broad categories of bear safety messaging: (1) general information; (2) how to avoid bear encounters; (3) how to defuse encounters; and (4) how to survive attacks. First, food and garbage should be secured to avoid attracting bears. People should move in bear areas in

246

groups of more than two and try to group together if a bear is encountered. Carrying bear spray is highly recommended. Additional specific precautions should be taken in urban areas and their proximities (Bombieri et al. 2018a). In such environments, increased attention to dogs and improved management of attractants (i.e. avoid leaving garbage, pet food, and bird feeders outside houses) would likely reduce the probability of risky encounters with this species. Moreover, to reduce the occurrence of predatory attacks, particular attention should be paid with children, who need to be constantly and strictly supervised by adults (Garrote et al. 2017; Penteriani et al. 2017). Finally, in case of an attack, being able to recognize the motivation behind it may be crucial in determining the attack outcome. That is, in the case of a predatory attack, one should try to aggressively deter the bear and fight back in any possible way. Instead, if the attack is defensive, one should be passive and adopt a defensive posture by lying face-down and protecting the neck with the hands.

Brown Bear Based on findings by Smith and Herrero (2018), actions that avoided brown bear encounters include being wary during their most active time of day (midday through evening), avoiding areas with poor visibility when possible, making noise to alert bears to one’s presence, and hiking in a party (2 people). These safety measures have also been recommended for forest users in Scandinavia (Støen et al. 2018), where bears reside in rugged terrain (Nellemann et al. 2007), rest in dense vegetation during daytime to avoid people (Ordiz et al. 2011), and generally flee when encountered (Moen et al. 2012). Of particular note, brown bear attacks, including fatalities, have been documented in every month of the year, underscoring the fact that one must employ bear avoidance practices year-round. Similar to advice highlighted for black bear encounters, carrying bear spray and grouping together when encountering a bear is recommended to avoid being attacked. In Scandinavia, awareness and education efforts, especially among hunters, have also been highlighted to ensure human safety (Støen et al. 2018). Because brown bear attacks are mostly defensive, if attacked one must be passive and adopt a defensive posture. Importantly, Smith and Herrero (2018) reported that, when rescuers came to the aid of the victims, the mauling ended 91% of the time, highlighting the importance of coming to the aid of the victim. Although dogs might help terminate an attack, one should make sure they obey commands and, preferably, keep them on a leash to avoid them disturbing bears and therefore triggering an aggressive reaction toward the dog and its owner.

Polar Bear Wilder et al. (2017) cautioned that, as sea ice continues to shrink, human–polar bear conflict can be expected to rise.

Patterns of Bear Attacks

Importantly, Fleck and Herrero (1988) observed that the outcome of human–polar bear conflict was most often a dead bear and much more rarely an injured or dead human. Therefore, improved conflict investigation is needed to collect accurate and relevant data and communicate accurate bear safety messages and mitigation strategies to the public. With better information, people can take proactive measures in polar bear habitat to ensure their safety and prevent conflicts with polar bears. To conclude, although rare, bear attacks on humans do occur within the whole range of bear species and undermine bear conservation efforts (Røskaft et al. 2007). Different bear species showed differing attack patterns, and although some bear species, such as the Andean bear, the sun bear, and the giant panda have never or rarely been reported to be involved in such incidents, the other five species of bears may locally represent a more serious threat to human safety. Therefore, it is of the utmost importance for bear conservation worldwide to reduce such conflicts by developing and implementing effective strategies based on both species-specific characteristics and the local socioeconomic context. In developed countries, where most attacks occur on people involved in recreational activities in bear areas, conflicts can be decreased through education and outreach (i.e. providing accurate bear safety messaging to the public). For instance, recent efforts in

References Acharya, K. P., Paudel, P. K., Neupane, P. R. & Köhl, M. (2016). Human–wildlife conflicts in Nepal: patterns of human fatalities and injuries caused by large mammals. PLoS ONE 11: 1–18. Akhtar, N. (2006). Human–sloth bear conflict: a threat to sloth bear conservation. International Bear News 15: 15–17. Akiyama, G., Kuwahara, H., Asahi, R., Tosa, R. & Yokota, H. (2017). Prompt procedures have a great impact on the consequences of Asiatic black bear mauling. Journal of Nippon Medical School 84: 294–300. Ali, A., Waseem, M., Teng, M., et al. (2018). Human–Asiatic black bear (Ursus thibetanus) interactions in the Kaghan Valley, Pakistan. Ethology Ecology and Evolution 30: 399–415. Anderson, K. (1957). Man-eaters and jungle killers. London: George Allen and Unwin Ltd. Arimoto, I., Okamura, H., Koike, S., Yamazaki, K. & Kaji, K. (2014). Behavior and habitat of Asiatic black bear (Ursus thibetanus) inhabiting near settlements. Honyuruikagaku 54: 19–31.

Yosemite National Park have demonstrated that with effective education, outreach, and appropriate penalties for the uncompliant, bear conflicts can be dramatically decreased (down by 99% since 1998), benefitting both people and bears (National Park Service 2019). A different approach should be employed in developing countries, where people involved in attacks are mostly local people who must enter bear areas for their daily work- and subsistence-related activities. Here, education on how to avoid bear encounters and attacks might not be sufficient alone and must be combined with improved facilities and financial support for local communities living in bear areas.

Acknowledgments

V.P. and A.O. were financially supported by the Excellence Project CGL2017–82782-P financed by the Spanish Ministry of Science, Innovation and Universities, the Agencia Estatal de Investigación (AEI), and the Fondo Europeo de Desarrollo Regional (FEDER, EU). G.B. was financially supported by a collaboration contract with MUSE – the Museo delle Scienze of Trento (Italy). M.M.D. was financially supported by the Spanish Ramon y Cajal grant RYC-2014-16263. J.M.F. and P.J.G. were funded by the Portuguese Foundation for Science and Technology (IF/00728/2013 and SFRH/BD/130527/2017).

Bargali, H. S., Akhtar, N. & Chauhan, N. P. S. (2005). Characteristics of sloth bear attacks and human casualties in North Bilaspur Forest Division, Chhattisgarh, India. Ursus 16: 263–267.

Charoo, S. A., Sharma, L. K. & Sathyakumar, S. (2011). Asiatic black bear–human interactions around Dachigam National Park, Kashmir, India. Ursus 22: 106–113.

Bombieri, G., Delgado, M. del M., Russo, L. F., et al. (2018a). Patterns of wild carnivore attacks on humans in urban areas. Scientific Reports 8: 17728.

Clark, D. (2003). Polar bear–human interactions in Canadian national parks, 1986–2000. Ursus 14: 65–71.

Bombieri, G., Nanni, V., Delgado, M. del M., et al. (2018b). Content analysis of media reports on predator attacks on humans: toward an understanding of human risk perception and predator acceptance. BioScience 68: 577–584. Bombieri, G., Naves, J., Penteriani, V., et al. (2019). Brown bear attacks on humans: a worldwide perspective. Scientific Reports 9: 8573. https://doi.org/10.1038/s41598019-44341-w Brower, C. (1942). Fifty years below zero: A lifetime of adventure in the Far North. New York, NY: Dodd, Mead and Company. Burton, R. G. (1856). A book of man eaters (first edition 1931). Delhi: Mittal Publications. Chapron, G., Kaczensky, P., Linnell, J.D.C., et al. (2014). Recovery of large carnivores in Europe’s modern human-dominated landscapes. Science 346: 1517–1519.

Conover, M.R. (2008). Why are so many people attacked by predators? Human– Wildlife Interactions 2: 139–140. Debata, S., Swain, K. K., Sahu, H. K. & Palei, H. S. (2017). Human–sloth bear conflict in a human-dominated landscape of northern Odisha, India. Ursus 27: 90–98. Dhamorikar, A. H., Mehta, P., Bargali, H. & Gore, K. (2017). Characteristics of human– sloth bear (Melursus ursinus) encounters and the resulting human casualties in the Kanha–Pench corridor, Madhya Pradesh, India. PLoS ONE 12: 1–18. Dharaiya, N., Bargali, H. S. & Sharp, T. (2016). Melursus ursinus. The IUCN Red List of Threatened Species 2016: e.T13143A45033815 (WWW document). http://dx.doi.org/10.2305/IUCN.UK .2016-3.RLTS.T13143A45033815.en Fenton, L. L. (1909). Tiger attacking a bear. Bombay Natural History Society 979.

247

Human–Bear Coexistence

Fleck, S. & Herrero, S. (1988). Polar bear– human conflicts. Contract report for Parks Canada and GNWT, contract 502/85/23. Garcia, K. C., Joshi, H. M. & Dharaiya, N. (2016). Assessment of human–sloth bear conflicts in North Gujarat, India. Ursus 27: 5–10. Garrote, P. J., Delgado, M. del M., LópezBao, J. V., et al. (2017). Individual attributes and party affect large carnivore attacks on humans. European Journal of Wildlife Research 63: 80. Herrero, S. (1972). Aspects of evolution and adaptation in American black bears (Ursus americanus Pallas) and brown and grizzly bears (U. arctos Linné.) of North America. Bears: Their Biology and Management 2: 221. Herrero, S. (2018). Bear attacks: Their causes and avoidance, 3rd edition. New York, NY: Lyons & Burford. Herrero, S. & Higgins, A. (2003). Human injuries inflicted by bears in Alberta: 1960–98. Ursus (Knoxville) 14: 44–54. Herrero, S., Higgins, A., Cardoza, J. E., Hajduk, L. I. & Smith, T. S. (2011). Fatal attacks by American black bear on people: 1900–2009. Journal of Wildlife Management 75: 596–603. Hwang, M. H. & Garshelis, DL . (2007). Activity patterns of Asiatic black bears (Ursus thibetanus) in the Central Mountains of Taiwan. Journal of Zoology 271: 203–209. Jamtsho, Y. & Wangchuk, S. (2016). Assessing patterns of human–Asiatic black bear interaction in and around Wangchuck Centennial National Park, Bhutan. Global Ecology and Conservation 8: 183–189. Jangid, K. A. & Sharma, K. R. (2018). How locals characterize the causes of sloth bear attacks in Jawai, Rajasthan. International Bear News 27: 11–12. Japan Bear Network. (2006). Understanding Asian bears to secure their future. Ibaraki: Japan Bear Network. Japan Bear Network. (2011). Report on statistics of the bear caused human injuries (in Japanese). Ibaraki: Japan Bear Network. Kozakai, C., Yamazaki, K., Nemoto, Y., et al. (2011). Effect of mast production on home range use of Japanese black bears. Journal of Wildlife Management 75: 867–875.

248

Kurt, F. & Jayasuriya, A. (1968). Notes on a dead bear. Loris 11: 182–183. Lal Moten, T., Bhat, T. A., Gulzar, A., Mir, A. & Mir, F. (2017). Causalities of human wildlife conflict in Kashmir valley, India; a neglected form of trauma: our 10 year study. International Journal of Research in Medical Sciences 55: 1898–1902. Laurie, A. & Seidensticker, J. (1977). Behavioural ecology of the sloth bear (Melursus ursinus). Journal of Zoology 182: 187–204. Littledale, H. (1889). Bears being eaten by tigers. Journal of Bombay Natural History Society 4: 316. Mardaraj, P. C. (2015). Identifying key issues for the conservation of sloth bear (Melursus ursinus) in Rajnilgiri, Odisha, Eastern India. The Rufford Foundation, United Kingdom. Mardaraj, P. & Dutta, S.K. (2011). Human– sloth bear conflict in Balasore Forest Division, Eastern India. Saarbrücken, Germany: LAP Lambert Academic Publishing GmbH & Co. KG. Moen, G. K., Støen, O. G., Sahlén, V. & Swenson, J. E. (2012). Behaviour of solitary adult Scandinavian brown bears (Ursus arctos) when approached by humans on foot. PLoS ONE 7: e31699. Nabi, D. G., Tak, S. R., Kangoo, K. A. & Halwai, M. A. (2009a). Comparison of injury pattern in victims of bear (Ursus thibetanus) and leopard (Panthera pardus) attacks. A study from a tertiary care center in Kashmir. European Journal of Trauma and Emergency Surgery 35: 153–158.

attacks in Japan: a description of seven cases. Forensic Science International 286: e14–e19. Penteriani, V., Delgado, M. del M., Pinchera, F., et al. (2016). Human behaviour can trigger large carnivore attacks in developed countries. Scientific Reports 6: 20552. Penteriani, V., Bombieri, G., Fedriani, J. M., et al. (2017). Humans as prey: coping with large carnivore attacks using a predator–prey interaction perspective. Human–Wildlife Interactions 11: 192–207. Rajpurohit, K. S. & Krausman, P. R. (2000). Human–sloth bear conflicts in Madhya Pradesh, India. Wildlife Society Bulletin 28: 393–399. Ramachandran, A. (2009, November 11). Frozen with fear: stranded teen v polar bears. The Sydney Morning Herald (WWW document). Available from www .smh.com.au/world/frozen-with-fearstranded-teen-v–polar-bears-20091111i9fs.html Rasool, A., Wani, A. H., Darzi, M. A., et al. (2010). Incidence and pattern of bear maul injuries in Kashmir. Injury 41: 116–119. Ratnayeke, S., Van Manen, F. T., Pieris, R. & Pragash, V. S. J. (2014). Challenges of large carnivore conservation: sloth bear attacks in Sri Lanka. Human Ecology 42: 467–479. Røskaft, E., Händel, B., Bjerke, T. & Kaltenborn, B. P. (2007). Human attitudes towards large carnivores in Norway. Wildlife Biology 13: 172–185.

Nabi, D. G., Tak, S. R., Kangoo, K. A. & Halwai, M. A. (2009b). Increasing incidence of injuries and fatalities inflicted by wild animals in Kashmir. Injury 40: 87–89.

Sethy, J. & Chauhan, N.S. (2013). Human– sun bears conflict in Mizoram, North East India: impact and conservation management. International Journal of Conservation Science 4: 317–328.

National Park Service. (2019). Bear facts (WWW document). Available from www .nps.gov/yose/planyourvisit/bearfacts.htm

Sethy, J. & Chauhan, N. S. (2016). Status and distribution of Malayan sun bear in Namdapha Tiger Reserve, Arunachal Pradesh, India. International Journal of Conservation Science 7: 533–552.

Nellemann, C., Støen, O. G., Kindberg, J., et al. (2007). Terrain use by an expanding brown bear population in relation to age, recreational resorts and human settlements. Biological Conservation 138: 157–165. Ordiz, A., Støen, O. G., Delibes, M. & Swenson, J. E. (2011). Predators or prey? Spatio-temporal discrimination of human-derived risk by brown bears. Oecologia 166: 59–67. Oshima, T., Ohtani, M. & Mimasaka, S. (2018). Injury patterns of fatal bear

Shah, A., Mir, B., Ahmad, I., et al. (2010). Pattern of bear maul maxillofacial injuries in Kashmir. National Journal of Maxillofacial Surgery 1: 96. Sharp, T. & Sonone, S. D. (2011). Sloth bear attacks: causes and consequences. International Bear Newsletter 20: 14–17. Sharp, T. R., Swaminathan, S., Arun, A. S., et al. (2017). Sloth bear attack behavior

Patterns of Bear Attacks

and a behavioral approach to safety. Final report to International Association for Bear Research and Management. Silwal, T., Kolejka, J., Bhatta, B. P., et al. (2017). When, where and whom: assessing wildlife attacks on people in Chitwan National Park, Nepal. Oryx 51: 370–377. Singh, N., Sonone, S., Rot, J. & Dharaiya, N. A. (2017). An unusual attractant spurs sloth bear break-ins in Maharashtra, India 26: 20–21. Smith, T. S. & Herrero, S. (2018). Human–bear conflict in Alaska: 1880–2015. Wildlife Society Bulletin 42: 254–263.

Støen, O. G., Ordiz, A., Sahlén, V., et al. (2018). Brown bear (Ursus arctos) attacks resulting in human casualties in Scandinavia 1977–2016; management implications and recommendations. PLoS ONE 13: e0196876. Swenson, J. E. (1999). Does hunting affect the behavior of brown bears in Eurasia? Ursus 11: 157–162. Tak, S. R., Nabi, D. G., Halwai, M. A. & Mir, B. A. (2009). Injuries from bear (Ursus thibetanus) attacks in Kashmir. Turkish Journal of Trauma & Emergency Surgery 15: 130–134. Wikipedia. (2019). Willem Barentsz (WWW socument). Available from https://en .wikipedia.org/wiki/Willem_Barentsz

Smith, T. S., Herrero, S., Debruyn, T. D. & Wilder, J. M. (2008). Efficacy of bear deterrent spray in Alaska. Journal of Wildlife Management 72: 640–645.

Wilder, J. M., Vongraven, D., Atwood, T., et al. (2017). Polar bear attacks on humans: implications of a changing climate. Wildlife Society Bulletin 41: 537–547.

Smith, T. S., Herrero, S., Layton, C. S., Larsen, R. T. & Johnson, K. R. (2012). Efficacy of firearms for bear deterrence in Alaska. Journal of Wildlife Management 76: 1021–1027.

Wilson, E. O. (1984). Biophilia. Cambridge, MA: Harvard University Press. Windler, A. (2014). Views towards the sun bear and frames on the human–sun bear

conflict of local people in West Sumatra, Indonesia. Wageningen: Forest and Nature Conservation, Wageningen University & Research. Yamazaki, K. (2004). Recent bear–human conflicts in Japan. International Bear News 13: 16–17. Yamazaki, K. (2010). Ursus thibetanus. In: Ohdachi, S. D., Ishibashi, Y., Iwasa, M. A., Fukui, D. & Saitoh, T. (Eds.), The wild mammals of Japan (pp. 243–245). Kyoto: Shoukado Book Seller. Yamazaki, K. (2017). Consecutive fatal attacks by Asiatic black bear on humans in Northern Japan. International Bear News 26: 16–17. Yamazaki, K. & Sato, Y. (2014). Countrywide range mapping of Asiatic black bears reveals increasing range in Japan. International Bear News 23: 18–19. Yamazaki, K., Furubayashi, K., Kasai, S., et al. (2008). A preliminary evaluation of activity-sensing GPS collars for estimating daily activity patterns of Japanese black bears. Ursus 19: 154–161.

249

Chapter

18

Effects of Human Disturbance on Brown Bear Behavior Ole-Gunnar Støen, Andre´s Ordiz, Marcus Elfstro¨m, Anne G. Hertel, Veronica Sahle´n, Jonas Kindberg, and Jon E. Swenson

Introduction Human persecution is the main cause of mortality for large carnivores in human-dominated landscapes, often due to conflicts with people, because they prey on livestock or wild game or harm people (Kaczensky 1997; Woodroffe & Ginsberg 1998; Røskaft et al. 2003; Nielsen et al. 2005; see also Chapters 15 and 17). This mortality affects population dynamics of large carnivores, leading to a strong association between high human density and the loss of regional carnivore populations (Woodroffe 2000). Human disturbance may be regarded as a predation risk, forcing large carnivores to behave as prey and consequently avoid humans to increase their survival (Lima 1998; Beale & Monaghan 2004; Ordiz et al. 2011). In the “landscape of fear” concept, this predation risk forces individual carnivores to alter their use of the landscape and invest more time in antipredator behavior (vigilance) and less time in profitable activities (foraging), which may affect reproduction (Altendorf et al. 2001; Laundré et al. 2001). Large carnivores in humanized landscapes may alter their behavior and avoid humans by being nocturnal and elusive (Basille et al. 2009; Knoppf et al. 2014; Ordiz et al. 2014). In some areas, favorable management policies allowed large carnivore populations to increase and even expand into more human-dominated landscapes (Chapron et al. 2014). Given the current scenario of human population growth, human expansion into wilderness areas, and increasing habitat loss (Hoekstra et al. 2005), the future survival of large carnivores may depend on their ability to adapt to live in human-dominated landscapes (Linnell et al. 2001). Bears are disturbed in many ways, either directly when they encounter humans or indirectly by changing their behavior and way of life to avoid humans, human activity, and infrastructure. Here we summarize research on how brown bears normally react when encountering humans, what a human encounter may entail for a bear, and whether bears become habituated or change their behavior toward humans with increased exposure. Based on this, we also discuss: (a) how our knowledge of brown bear behavior may help people to deal with their fear of bears, and not limit their use of outdoor areas with bears; (b) how human presence, activity, and infrastructure have an indirect effect on bears, that is how bears change their movement pattern, use of terrain and vegetation, and

250

daily activity pattern to avoid humans; (c) how human disturbance influences foraging and denning, which is crucial for brown bear growth and reproduction; and (d) apparent differences between continents in brown bear behavior toward humans and whether this may have an evolutionary cause.

Brown Bear Behavior When Encountering Humans Brown bears can avoid humans in time and space by being mostly nocturnal and residing far from people (Nellemann et al. 2007; Martin et al. 2010). However, bears cannot always avoid encounters with humans, especially in areas close to human settlements, areas with much human activity, or even in back-country areas, where hikers may unintentionally encounter bears. Some of these encounters may result in dangerous situations, where both humans and bears can be harmed (see Chapter 17). However, the majority of such encounters are uneventful, with the bears’ presence going completely unnoticed (Moen et al. 2012). Several studies have aimed to understand how bears normally react during encounters with humans (e.g. Swenson et al. 1999). The purpose of such studies is also to be able to inform the public and management authorities about normal bear behavior, how to behave when encountering bears, and how to prevent dangerous situations from occurring. The first studies in Scandinavia, which relied on direct observations and analyzed the outcome of encounters between bears and people, concluded that bears were not aggressive toward humans and seldom attacked, and that in the cases the bear had charged at the human, these were bluff charges to scare the human away (Swenson et al. 1999). These earlier studies were by design limited to studying how bears reacted during encounters when they were seen by the observers, and only for the time that they were being observed. Therefore, they were unable to study the bears’ behavior during encounters when they were out of sight, e.g. when in dense vegetation. Further, as bears are elusive animals, these studies can only be expected to cover a small fraction of bear–human encounters, as most encounters are likely to go unnoticed by humans. The use of geographical positioning systems (GPS) technology in wildlife research has enabled

Effects of Human Disturbance

researchers to study animal behavior remotely and in greater detail than in previous studies. This technology provides the opportunity to study how bears behave during encounters where they are not observed, as well as information on how often bear–human encounters occur without being detected by humans. Such studies have been carried out in the Scandinavian brown bear population, using experimental approaches that simulated the behavior of hikers in order to study how single bears (Moen et al. 2012) and females with cubs (Sahlén et al. 2015a) respond to encounters when they are aware that humans are approaching. The studies revealed that the absolute majority of encounters go unnoticed by humans (Moen et al. 2012; Ordiz et al. 2013). When bears are aware that humans are approaching, they either leave the area or hide until the humans have passed. As a result, observers only saw the bears in 15% (single bears) or 26% (females with cubs) of approaches, despite knowing exactly where the bear was located. The difference in detectability between single bears and females with cubs appears to be due to females with cubs preferring more open habitat, probably as a strategy to avoid adult infanticidal males (Sahlén et al. 2015b). Most bears, whether active or resting in a day bed, left their location when humans approached. Whether they left depended on how close the humans passed the bears (Moen et al. 2012). The distance from the observer at which bears decided to leave, the flight initiation distance (FID), was longer for active single bears (average 115 m) that were already on their feet than for passive single bears (average 69 m) lying in their resting site. Younger single bears left at longer distances than older single bears, probably because they were less experienced with encountering people. The FID was also longer when the bears were in more open habitat, with less cover to hide in. FID did not differ between single bears and females with cubs, but a greater percentage of females with cubs left as a response to an encounter with humans (Sahlén et al. 2015a). Following the encounter, bears that left as a response to the disturbance moved on average 1.2 km before settling into a new resting site. Females with cubs displayed stronger immediate reactions, by moving farther before settling into a new resting site (Sahlén et al. 2015a). The reactions by bears to encounters with humans reflected on their daily activity patterns, and this effect went beyond the same day of the approach. Bears were more active, i.e. they moved larger distances during nocturnal hours, and less during the day, for several days following the encounter (Ordiz et al. 2013 and Figure 18.1). The bears reduced their movement by 10% and 11% during daytime the first two days following an experimental encounter, respectively, but showed a similar increase in activity during the darkest hours of the night when they normally rest, probably to compensate for disrupted feeding during the day (Ordiz et al. 2013). The closer the observers had passed and the denser the vegetation at the encounter location, the stronger the disturbance effect. This indicates that bears react more strongly when disturbed at short distances in highly concealed sites, where bears hide from people during their day rest (Ordiz et al. 2011).

Brown bear reactions to encountering humans seem to be consistent between populations and varying human density. In Fennoscandia, a comparative study showed that the proportion of bears that fled when encountered by humans and the distance from which they initiated their flight were similar between bears in Sweden and Finland, where bears lived in areas with a higher human density (Moen et al. 2019). Similarly, in Sweden, the density of humans and roads within the home range of the bears, used as a proxy for probable encounter rates with humans, did not influence the flight responses of the bears. This consistency may be explained by a similar history in Fennoscandia, with recent population recovery after near extirpation of the brown bear population (Pulliainen 1983; Swenson et al. 1995, 2017; Kopatz et al. 2014), where bears are managed as a game species, and have experienced long-term persecution, which may make bears elusive in Fennoscandia (Ordiz et al. 2011; Zedrosser et al. 2011). However, Fennoscandia has among the lowest human population densities across the distribution range of brown bears, at least in Europe (Swenson et al. 2000), so similar studies might be needed on bears in areas with higher human density to fully understand the normal behavior of bears when encountered by humans (Moen et al. 2019).

Brown Bear Habituation and Sensitization Habituation is a process that causes decreasing responsiveness to a repeated stimulus (Thompson 2009), whereas sensitization is an increased responsiveness to a repeated stimulus (Blumstein 2016). Although brown bears generally avoid humans, some individuals can habituate to people in close proximity. In North America, brown bear habituation to hikers and anglers, for instance, has been documented in different ecosystems and national parks (e.g. McLellan & Shackleton 1989; Herrero et al. 2005). Habituated bears face higher mortality risk, which in turn can have a negative effect on bear population dynamics. For example, habituated bears near roads or railways are more likely to be killed by vehicles or trains and to be removed legally and illegally (Herrero et al. 2005). In Yellowstone National Park, USA, bear habituation to people and sensitization to human-derived food (i.e. food conditioning) resulted in human fatalities a few decades ago, which caused the removal of many bears and a population decline (e.g. Penteriani et al. 2018). In Europe, bears are also generally wary of humans, but this can change and requires that human activities that can alter bear behavior be actively managed to prevent habituation (Penteriani et al. 2018). Subadult bears may approach human habitation in order to avoid adult conspecifics and/or because of a lack of experience, i.e. naivety, which, after exposure to people, may result in human habituation or food conditioning (Elfström et al. 2014b). Therefore, subadults may be more vulnerable to habituation and food conditioning than older, experienced bears that generally avoid people.

251

Human–Bear Coexistence

(A)

Figure 18.1 Estimated time effect, every 30 min during the 24 h day, on the daily activity pattern of brown bears in central Sweden during the week before the experimental approach. Bears typically had a main resting period during midday and a second resting period around midnight (A). The daily activity pattern of bears on the day of the approach, showing the initial escape after the encounter, followed by a reduction in movement. Vertical lines illustrate when most approaches were conducted (B). Bears became more nocturnal in the following days after the approach, compared to the daily pattern in the week before the approach. The vertical lines illustrate when most approaches were conducted (C). In all panels, curves represent the mean of the distance travelled and the 95% credible intervals. Figures from Ordiz et al. (2013). (© 2013 The Authors. Journal of Applied Ecology © 2013 British Ecological Society.)

16

Square root meters

14

12

10

8

6 0

1

2

3

4

5

6

7

8

9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 Time

(B) 16

Square root meters

14

12

10

8

6 0

1

2

3

4

5 6

7

8

9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 Time

(C)

3

+26%

Difference

2

+11%

+8%

1 0 –1 –2

–10% –10% Day 0

252

Day 1

–11%

Day 2

Day 3 Day 4 Time

Day 5

Day 6

Day 7

Effects of Human Disturbance

A recent study in Scandinavia (Ordiz et al. 2019) performed repeated approaches to GPS-collared brown bears to document the potential habituation or sensitization to repeated, consistently similar stimuli. Bears were approached up to eight consecutive times in a few weeks, with no clear change in the bears’ general response, i.e. avoiding people, in terms of flight initiation distance and proportion of bears that stayed or moved from their position before the approaches started (Ordiz et al. 2019). This suggests that, at the level of experimentally induced disturbance, bears did not show habituation or sensitization to the repeated stimuli. Interestingly, there is a remarkable scarcity of studies on wildlife habituation or sensitization that have analyzed the response of radio-collared animals, which is essential to determine if specific individuals habituate or not and, consequently, to inform managers (Blumstein 2016). Besides the observation of clearly habituated individuals that tolerate human presence, subtle, physiological and nonvisible changes that incur energetic costs can occur in response to humans (Herrero et al. 2005; Støen et al. 2015). It has been suggested that genetic differences between habituated and nonhabituated individuals may exist within a bear population (Cotovelea et al. 2015). Nevertheless, bear habituation may reduce other negative effects of human activity on large carnivores, such as the total avoidance of some areas; therefore, further studies are needed to fully understand the potential population-level consequences of habituation (Wheat & Wilmers 2016). Population-level effects may be different depending on the bear population size and conservation status, degree of human encroachment, and exposure of individual bears to human activities.

Human Effects on Brown Bear Activity Rhythms, Movement Patterns, and Habitat Use Human persecution has eradicated brown bears from many places and, where bears still exist, human activities have reduced bear habitat quality and the size of many bear populations (e.g. Suring et al. 1998; Proctor et al. 2010; Ordiz et al. 2017). Therefore, human activities have effects on bear population trends and, in human-dominated landscapes, bears change or adapt their behavior to deal with humans. Bear adaptations include changes in movement patterns, terrain use, and daily activity patterns, which have been documented at different spatial and temporal scales (e.g. Ordiz et al. 2017). Across their widespread range, brown bears select for rugged terrain, where human access is limited (Apps et al. 2004; Nellemann et al. 2007; Martin et al. 2010). At finer scales, bears select for dense vegetation cover and avoid areas near human settlements and paths when choosing resting sites, especially when human outdoor activity is greatest (Ordiz et al. 2011; Cristescu et al. 2013; Figures 18.2 and Figure 18.3). Human activity also alters the optimal daily activity patterns of

Figure 18.2 A typical resting site of a brown bear in Sweden. The bear rested among the vegetation-covered rocks with good cover from surrounding trees and larger rocks. Hairs of the bears are often found in the resting bed. (Photo by O.-G. Støen.)

brown bears. Disturbance can exclude bears completely from an area or reduce its time of use (Suring et al. 1998). Generally, bears are more active during daytime in areas with less human activity, and tend to be nocturnal in more humanized landscapes (e.g. Machutchon et al. 1998; Olson et al. 1998; Kaczensky et al. 2006; Martin et al. 2010). Reduced daytime activity is particularly clear after bears encounter people (McLellan & Shackleton 1989; Moen et al. 2012). Where bears are hunted, they also become more nocturnal when hunting seasons start (Ordiz et al. 2012). Bears avoid human infrastructure, e.g. roads (Mace et al. 1996; Elfström et al. 2008) or are more nocturnal in areas with higher road density (Ordiz et al. 2014). Thus, limiting access or closing roads that no longer are necessary, e.g. after logging, is advisable to reduce disturbance (Ordiz et al. 2014; Sorensen et al. 2015). Adult bears, especially males, avoid human settlements, cabins, and ski resorts, whereas areas closer to people are used by younger individuals (Mueller et al. 2004; Nellemann et al. 2007; Elfström et al. 2014a, 2014b; Sorensen et al.

253

Human–Bear Coexistence

Figure 18.3 Across their widespread range, brown bears select for rugged terrain, where human access is limited. (Photo by V. Penteriani.)

2015). In turn, bears face higher mortality risk near people and human settlements (Steyaert et al. 2016). In summary, bears react to all types of human activities and bear reactions are visible at all spatial and temporal scales (Ordiz et al. 2017). In bear populations living in humandominated landscapes under long-term protection, bears continue to avoid people, yet differences in bear behavior in relation to human infrastructure may not be as clear as in hunted populations, showing a general pattern of human avoidance that does not seem to change in relation with distance to human infrastructure (Zarzo-Arias et al. 2018).

Human Effects on Brown Bear Foraging In most animals, effective and sufficient foraging is the premise for successful reproduction. Bears give birth in the winter den and are therefore particularly dependent on finding enough food during summer prior to hibernation. Humans affect food availability and bear foraging efficiency in multiple ways. On one hand humans may intentionally or unintentionally provision bears with food. Although bears near human habitation primarily seek shelter, this may or may not result in foraging on anthropogenic foods (Elfström et al. 2014a, 2014b, 2014c). Human infrastructure provides easy access to anthropogenic food resources, like trash, and bears can readily use these resources (Bojarska & Selva 2012; Johnson et al. 2015). Ungulate hunting facilitates highly nutritious food when hunters leave carrion or gut piles of game that bears scavenge on. Further, in six of nine brown bear populations in Europe, cooccurring ungulates are given supplementary feed during winter and, despite not being the target species, bears regularly use these feeding sites (Selva et al. 2017). Elsewhere, intentional feeding of bears by humans is common. Several European countries, including Slovenia, Croatia, Serbia, and Romania,

254

use diversionary feeding at bait sites to lure bears away from human settlements (Kavcic et al. 2013). Bait sites may likewise be used to attract bears either for hunting or for bear-viewing tourism (Bischof et al. 2008; Penteriani et al. 2017). Contrary to provisioning bears, human activity, presence, and infrastructure have also been shown to constrain bear foraging activities. Brown bears in Scandinavia have been shown to use less-productive foraging areas when avoiding humans temporally (Hertel et al. 2016; Figure 18.4) or spatially (Lodberg-Holm et al. 2019). Salmon streams in Alaska attract bears in high numbers and hence are popular for bear-viewing tourism. The presence of tourists, however, can displace bears spatiotemporally from salmon streams and therefore hamper food intake (Rode et al. 2006). Lastly, probably the most influential way humans alter food availability for bears is the transformation of land-use by mining, deforestation, or fire management (Nielsen et al. 2004; Cristescu et al. 2016). Tree or topsoil removal changes nutrient and light availability and hence fundamentally alters predominant plant communities on which bears forage. The pathways through which humans affect bear foraging depend largely on local wildlife and habitat management practices and are highly ecosystemspecific. Cascading effects can be trait-mediated indirect effects, where an antipredation strategy by a prey species influences species in other trophic levels (Abrams 1995). One example is the brown bear in Scandinavia, which practise antipredation strategies in relation to humans (Ordiz et al. 2011) and at the same time is a major predator on red wood ants (Formica rufa). Red wood ants are a keystone species in the boreal forest, influencing other species in the ecosystem (Haemig 1992, 1994; Kilpelainen et al. 2009; Stockan & Robinson 2016). Through predation and disturbance of anthills, brown bears reduce the abundance of ants (Moen 2018). Because brown

Effects of Human Disturbance

0.2

0.75

0.15 Morning

Afternoon

0.5

0.1

0.25

0.05

50

Pre-hunting season Hunting season

40 Efficiency (berries/m2)

Pre-hunting season Hunting season Mortality risk

Relative mortality risk

Probability of foraging activity

1

30 20 10 0

0

0 22

4

10 Time of day

16

22

Morning

Afternoon

Figure 18.4 Effect of mortality risk on foraging activity (left panel) and foraging efficiency (right panel) during the first two weeks of the hunting season, as compared to last two weeks before the hunting season in south-central Sweden. Left: Foraging activity was measured as slow, meandering movements. Relative mortality risk was measured as the number of bears killed at each time of day in the study area during the hunting seasons 2006–2014 (total number of shot bears = 680). Mortality risk peaked in the morning hours after the main foraging activity peak but was low when foraging activity was high in the afternoon. Bears reduced foraging activity in the morning but not in the afternoon during the hunting season. Right: Foraging efficiency was defined as the richness of the foraging location and was measured by how many more berries were found at a bear foraging location relative to how many berries would be expected in the same habitat and time of year. Bears selected the best foraging location during their morning foraging bout prior to the hunting season but decreased after onset of the hunting season. Bears probably use denser vegetation with more vertical cover but less light availability for berry-producing shrubs. Foraging efficiency in the afternoon when mortality risk was low was the same for the pre-hunting and hunting seasons. Modified from Hertel et al. (2016).

bears avoid human activity, humans may modify the effect bears have on red wood ants, and mediate cascading effects in the boreal forest (Moen 2018).

Human Effects on Brown Bear Denning During the winter denning period, brown bears avoid human disturbance by usually choosing den sites in steep terrain with few roads which are therefore less accessible for people (Petram et al. 2004; Elfström et al. 2008; Sahlén et al. 2011). Adult males often choose den sites in more remote areas, suggesting that the more predation-vulnerable females and subadult males also avoid dominant adult males during denning (Elfström & Swenson 2009; Elfström et al. 2014b). The level of human activity elicits increased avoidance behavior when bears choose den sites. Settlements and roads with high traffic that also provide direct human access are avoided when bears choose den sites (Elfström et al. 2008). Brown bears that den closer to plowed roads are at a higher risk of abandoning their dens compared to bears denning in areas with lower road accessibility (Elfström & Swenson 2009). As a result, bears that den closer to human activity also choose more concealed den sites, i.e. with increased horizontal cover and terrain ruggedness (Sahlén et al. 2011). Bears seem to be more sensitive to disturbance when they are about to enter their den sites, and abandonment rates are higher during the onset of the denning

period (Sahlén et al. 2015a). Female brown bears that abandon their dens can experience lower reproductive success (Swenson et al. 1997). Therefore, females may tolerate greater levels of disturbance without abandoning their dens after giving birth. This suggests that roads and human activity have a negative impact and represent a disturbance for bears. After bears have entered their dens, the risk of human injuries during bear– human encounters increases due to a combination of factors, including reduced bear physiological activity that prevents them from escaping from approaching people (Sahlén et al. 2015b; Støen et al. 2018).

Human Effects on Brown Bears in an Evolutionary Perspective We have discussed above how brown bears show plasticity in their reactions to variations in human disturbance. It is a common perception that the grizzly bear of North America is more aggressive (Herrero 1985; Swenson et al. 1999; but see Bombieri et al. 2019) and more diurnally active (Roth 1983; Ordiz et al. 2014) than the European brown bear, although they are the same species. Many factors, such as human density, bear density, habitat productivity and openness, and persecution and hunting, might be enough to explain geographic differences in brown bear behavior and life-history traits. Recent research has highlighted the potential for long-term

255

Human–Bear Coexistence

evolutionary effects of directed selection during heavy human harvesting (Coltman et al. 2003; Festa-Bianchet 2003; Proaktor et al. 2007; Allendorf & Hard 2009; Darimot et al. 2009; Mysterud 2011). Brown bears were persecuted systematically for many centuries in Europe, but for a relatively short time (400 bears for the entire Pindos population (Karamanlidis et al. 2015). This is indicative of a significant population recovery of the bear population in Pindos and is consistent with large carnivore population recoveries throughout Europe (Chapron et al. 2014). Bear densities obtained in the four sampling areas in Pindos were the lowest recorded for the species in the

Dinara-Pindos population (Huber et al. 2008; Jerina et al. 2013) and are consistent with the assumption of a DinaraPindos population with a core population in the North and extending towards the South (Kaczensky et al. 2013).

Behavioral Adaptations The range, genetic, and demographic recovery of bears in Pindos has also likely been facilitated by specific behavioral adaptations that enable bears to survive in human-modified landscapes. As it occurred during the recovery of brown bears in Scandinavia (Kindberg et al. 2011), bears in Pindos have shown high levels of behavioral plasticity as a response to the rapidly changing and ecologically challenging environment they inhabit. For example, the activity of adult bears is mainly nocturnal, enabling them to avoid human disturbance (De Gabriel Hernando et al. 2020). Similarly, bear habitat preferences revealed efforts to avoid human disturbance. Also, habitat use had a clear circadian pattern: in general, proximity to human-related habitat features (e.g. intensive crops, naturalized crops, human settlements, road networks, and unpaved roads) was higher at night, whereas proximity to natural habitat features, such as forests, shrublands, high altitudes, and areas with rough terrain was higher during the day for all sex/age classes (De Gabriel Hernando et al. in review). Despite the population recovery exhibited by the brown bear population in Pindos, the fate of the species has not yet been totally secured. Several threats still compromise the recovery of bears despite legal and institutional protection. These threats include human-caused mortality, e.g. poaching (Mertzanis 2000; Mertzanis et al. 2009), vehicle collisions (Karamanlidis & Mertzanis 2004), retaliatory killing for damage to property (Karamanlidis et al. 2011), and habitat loss, fragmentation, and alteration at a range scale, e.g. construction of major highways (Karamanlidis et al. 2012a, 2012b).

What Can We Learn from Past and Present Histories of Small and Endangered Bear Populations in Human-Modified Landscapes? The diverse situation, from ongoing recovery to apparent stagnation, of some brown bear populations and subpopulations in human-modified landscapes demands an improvement of our knowledge on how this species is able to coexist with humans (Carter & Linnell 2016) and, in turn, what effects human presence and activity may potentially have on them. This is particularly important in areas with recovering but still endangered brown bear populations (Treves & Karanth 2003; Ordiz et al. 2013), which inhabit areas with suitable habitat surrounded by greater human encroachment, thus creating scenarios that may become ecological traps (Naves et al. 2003b; Penteriani et al. 2018a; Scharf & Fernández 2018). The case of the Cantabrian bear population is particularly noteworthy from a conservation perspective as humaninduced mortality appears to have decreased, the potential

267

Human–Bear Coexistence

for population recovery seems quite high, and the future of the species in the area is promising. However, this population is small and still highly vulnerable to the potential impacts of varying human activities and behaviors, as reflected by the slower demographic response of the eastern subpopulation, compared to that of the western one. A direct comparison of the genomes of Cantabrian and Apennine bears, with the latter apparently less capable of demographic growth despite similar levels of protection, could possibly clarify the role of genetic load. On the basis of the population trajectories and current status of the three European populations presented in this study, we believe that it is necessary to include more individual-based research (e.g. telemetry in the Cantabrian Mountains) in order to extend the available information essential for effective conservation and management, such as bear rhythms of activity, dispersal, home range behavior, and highresolution space use. Additionally, more monitoring and conservation efforts should be accorded to dispersing bears (Maiorano et al. 2017), especially in the case of females or family groups detected outside the core range. Indeed, the identification, management, and conservation of critical corridors that may allow range expansion and connectivity among core habitats should be particularly focused on the biologically most important segments of bear populations, that is, females and juveniles (Balbontín et al. 2005; Maiorano et al. 2017; Morini et al. 2017). The future of each one of these three small populations will also depend on the cooperation and management of local administrations, irrespective of which administration has primary jurisdiction at a local scale. Specific regulations and agency responsibilities may change, but bears will require trans-regional coordination in conservation and management policies (Penteriani et al. 2018b). Knowing the areas into which these populations are likely to expand would allow authorities and conservation organizations to focus information campaigns and pre-emptive/proactive damage control actions in these areas. Such proactive approaches are

References Balbontín, J., Penteriani, V. & Ferrer, M. (2005). Humans act against the natural process of breeder selection: a modern sickness for animal populations? Biodiversity and Conservation 14: 179–186. Benazzo, A., Trucchi, E., Cahill, J. A., et al. (2017). Survival and divergence in a small group: the extraordinary genomic history of the endangered Apennine brown bear stragglers. Proceedings of the National Academy of Sciences 114: E9589–E0597. Bischof, R., Swenson, J. E., Yoccoz, N. G., Mysterud, A. & Gimenez, O. (2009). The magnitude and selectivity of natural and multiple anthropogenic mortality causes

268

important for successful large carnivore conservation and management (Ripple & Beschta 2012). In addition, as brown bears disappeared from certain areas decades ago, local communities are no longer familiar with them. Thus, local information campaigns directed at both residents and other users, e.g. hunters and tourists, of areas of potential bear expansion represent a crucial strategy to reduce human–bear conflicts and promote coexistence. Finally, it is worthwhile to note here that, especially in the case of small, isolated, and/or endangered animal populations, the effects of climate change on trophic resources may considerably override conservation and management efforts performed at other levels, e.g. reduction of human–wildlife conflicts, threat of anthropogenic footprints and activities, poaching, and habitat fragmentation. Thus, conservation plans for species at higher trophic levels, such as brown bears, should take into account climate change vulnerability assessments of those plant communities that represent primary food resources and shelter for the target species (Penteriani et al. 2019). Thus, together with conservation actions aimed at maintaining bears in their historical and current ranges, we encourage conservation and management practices targeted at those areas potentially favorable to habitation by bears during the recovery process, taking into account the current context of climate change.

Acknowledgments We thank the Administrations of the Gobierno del Principado de Asturias and the Junta de Castilla y León for providing the bear database. V.P. and A.O. were financially supported by the Excellence Project CGL2017–82782-P financed by the Spanish Ministry of Science, Innovation and Universities, the Agencia Estatal de Investigación (AEI), and the Fondo Europeo de Desarrollo Regional (FEDER, EU). E.R. was supported by project CGL2017–83045-R AEI/FEDER-UE from the Spanish Agencia Estatal de Investigación, Ministerio de Ciencia, Innovación y Universidades.

in hunted brown bears. Journal of Animal Ecology 78: 656–665. Bonnet Lebrun, A.-S., Karamanlidis, A. A., De Gabriel Hernando, M., Renner, I. & Gimenez, O. (2020). Identifying priority conservation areas for recovering large carnivores using citizen science data. Animal Conservation 23: 83–93. Carter, N. H. & Linnell, J. D. C. (2016). Coadaptation is key to coexisting with large carnivores. Trends in Ecology and Evolution 31: 575–578. Chapron, G., Kaczensky, P., Linnell, J. D. C., et al. (2014). Recovery of large carnivores in Europe’s modern human-dominated landscapes. Science 346: 1517–1519.

Ciucci, P. & Boitani, L. (2008). The Apennine brown bear: a critical review of its status and conservation problems. Ursus 19: 130–145. Ciucci, P., Tosoni, E., Di Domenico, G. Quattrociocchi, F. & Boitani, L. (2014). Seasonal and annual variation in the food habits of Apennine brown bears, central Italy. Journal of Mammalogy 95: 572–586. Ciucci, P., Gervasi, V., Boitani, L., et al. (2015). Estimating abundance of the remnant Apennine brown bear population using multiple noninvasive genetic data sources. Journal of Mammalogy 96: 206–220.

Human-Modified Landscapes

Ciucci, P., Altea, T., Antonucci, A., et al. (2017). Distribution of the brown bear (Ursus arctos marsicanus) in the central Apennines, Italy, 2005–2014. Hystrix 28: 1–6.

Apennine brown bear population size based on hair-snag sampling and multiple data source mark–recapture Huggins models. Ursus 19: 105–121.

Clevenger, A. P., Purroy, F. J. & Pelton, M. R. (1992). Food habits of brown bears (Ursus arctos) in the Cantabrian Mountains, Spain. Journal of Mammalogy 73: 415–421.

Gervasi, V., Ciucci, P., Boulanger, J., Randi, E. & Boitani, L. (2012). A multiple data source approach to improve abundance estimates of small populations: the brown bear in the Apennines, Italy. Biological Conservation 152: 10–20.

Colangelo, P., Loy, A., Huber, D., et al. (2012). Cranial distinctiveness in the Apennine brown bear: genetic drift effect or ecophenotypic adaptation? Biological Journal of the Linnean Society 107: 15–26.

Gervasi, V., Boitani, L., Paetkau, D., et al. (2017). Estimating survival in the Apennine brown bear accounting for uncertainty in age classification. Population Ecology 59: 119–130.

Falcucci, A., Maiorano, L., Ciucci, P., Garton, E. O. & Boitani, L. (2008). Landcover change and the future of the Apennine brown bear: a perspective from the past. Journal of Mammalogy 89: 1502–1511.

Gonzalez, E. G., Blanco, J. C., Ballesteros, F., et al. (2016). Genetic and demographic recovery of an isolated population of brown bear Ursus arctos L., 1758. PeerJ 4: e1928.

Falcucci, A., Ciucci, P., Maiorano, L., Gentile, L. & Boitani, L. (2009). Assessing habitat quality for conservation using an integrated occurrence-mortality model. Journal of Applied Ecology 46: 600–609. FAPAS/FIEP. (2017). El Oso. Pirineos y Cordillera Cantábrica. Gijón, Spain: Gráficas Muñiz. Frosch, C., Dutsov, A., Zlatanova, D., et al. (2014). Noninvasive genetic assessment of brown bear population structure in Bulgarian mountain regions. Mammalian Biology 79: 268–276. De Gabriel Hernando, M., Karamanlidis, A. A., Grivas, K., et al. (2020). Reducing movement in Mediterranean landscapes: a case study of brown bears in Greece. Journal of Zoology. https://doi .org.10.1111/jzo.12768 De Gabriel Hernando, M., Karamanlidis, A. A., Grivas, K., et al. (in review). Habitat ecology informs conservation priorities of an endangered brown bear population in a human-dominated landscape. Journal of Nature Conservation. García, P., Lastra, J., Marquínez, J. & Nores, C. (2007). Detailed model of shelter areas for the Cantabrian brown bear. Ecological Informatics 2: 297–307.

Graban, J., Kisková, J., Pepich, P. & Rigg, R. (2013). Genetic analysis for geographic isolation comparison of brown bears living in the periphery of the Western Carpathians Mountains with bears living in other areas. Open Journal of Genetics 3: 174–182. Huber, Đ., Jakšić, Z., Frković, A., et al. (2008). Brown Bear Management Plan for the Republic of Croatia. Zagreb, Croatia: Ministry of Regional Development, Forestry and Water Management, Directorate for Hunting Ministry of Culture, Directorate for the Protection of Nature. Jerina, K., Jonozovič, M., Krofel, M. & Skrbinšek, T. (2013). Range and local population densities of brown bear Ursus arctos in Slovenia. European Journal of Wildlife Research 59: 1–9. Kaczensky, P., Chapron, G., von Arx, M., et al. (2013). Status, management and distribution of large carnivores – bear, lynx, wolf and wolverine – in Europe. Report to the EU Commission. Karamanlidis, A. A. & Mertzanis, G. (2004). Fatal bear–car collision. International Bear News 13: 24. Karamanlidis, A. A., Youlatos, D., Sgardelis, S. & Scouras, Z. (2007). Using sign at power poles to document presence of bears in Greece. Ursus 18: 54–61.

Gervasi, V. & Ciucci, P. (2018). Demographic projections of the Apennine brown bear population Ursus arctos marsicanus (Mammalia: Ursidae) under alternative management scenarios. European Zoological Journal 85: 243–253.

Karamanlidis, A. A., Krambokoukis, L. & Kantiros, D. (2008). Challenges and problems arising from the range expansion of brown bears in Greece. International Bear News 17.

Gervasi, V., Ciucci, P., Boulanger, J., et al. (2008). A preliminary estimate of the

Karamanlidis, A. A., Drosopoulou, E., de Gabriel Hernando, M., et al. (2010).

Non-invasive genetic studies of brown bears using power poles. European Journal of Wildlife Research 56: 693–702. Karamanlidis, A. A., Sanopoulos, A., Georgiadis, L. & Zedrosser, A. (2011). Structural and economic aspects of human–bear conflicts in Greece. Ursus 22: 141–151. Karamanlidis, A. A., Beecham, J., Bousbouras, D., et al. (2012a). Bear– vehicle collisions: a holistic approach to evaluating and mitigating the effects of this new conservation threat to brown bears in Greece. Presented at 21st International Conference on Bear Research and Management, New Delhi, India. Karamanlidis, A. A., Straka, M., Drosopoulou, E., et al. (2012b). Genetic diversity, structure, and size of an endangered brown bear population threatened by highway construction in the Pindos Mountains, Greece. European Journal of Wildlife Research 58: 511–522. Karamanlidis, A. A., de Gabriel Hernando, M., Krambokoukis, L. & Gimenez, O. (2015). Evidence of a large carnivore population recovery: counting bears in Greece. Journal for Nature Conservation 27: 10–17. Karamanlidis, A. A., Skrbinšek, T., de Gabriel Hernando, M., et al. (2018). History-driven population structure and assymetric gene flow in a recovering large carnivore at the rear-edge of its European range. Heredity 120: 168–182. Kindberg, J., Swenson, J. E., Ericsson, G. & Bellemain, E. (2011). Estimating population size and trends of the Swedish brown bear Ursus arctos population. Wildlife Biology 17: 114–123. Lamamy, C., Bombieri, G., Zarzo-Arias, A., González-Bernardo, E. & Penteriani, V. (2019). Can different habitat composition and structure help explain the different trends of Cantabrian brown bear subpopulations? Mammal Research 64: 559–567. Loy, A., Genovesi, P., Galfo, M., Jacobone, M. G. & Vigna Taglianti, A. (2008). Cranial morphometrics of the Apennine brown bear (Ursus arctos marsicanus) and preliminary notes on the relationships with other southern European populations. Italian Journal of Zoology 75: 67–75. Maiorano, L., Boitani, L., Chiaverini, L. & Ciucci, P. (2017). Uncertainties in the identification of potential dispersal

269

Human–Bear Coexistence

corridors: the importance of behaviour, sex, and algorithm. Basic and Applied Ecology 21: 66–75. Martin, J., Revilla, E., Quenette, P. Y., et al. (2012). Brown bear habitat suitability in the Pyrenees: transferability across sites and linking scales to make the most of scarce data. Journal of Applied Ecology 49: 621–631. Martínez Cano, I., González Taboada, F., Naves, J., Fernández-Gil, A. & Wiegand, T. (2016). Decline and recovery of a large carnivore: environmental change and long-term trends in an endangered brown bear population. Proceedings of the Royal Society B 283: 20161832. Mateo-Sánchez, M. C., Cushman, S. A. & Saura, S. (2014a). Scale dependence in habitat selection: the case of the endangered brown bear (Ursus arctos) in the Cantabrian Range (NW Spain). International Journal of Geographical Information Science 28: 1531–1546. Mateo-Sánchez, M. C., Cushman, S. A. & Saura, S. (2014b). Connecting endangered brown bear subpopulations in the Cantabrian Range (north-western Spain). Animal Conservation 17: 430–440. Mateo-Sánchez, M. C., Balkenhol, N., Cushman, S., et al. (2015a). A comparative framework to infer landscape effects on population genetic structure: are habitat suitability models effective in explaining gene flow? Landscape Ecology 30: 1405–1420. Mateo-Sánchez, M. C., Balkenhol, N., Cushman, S., et al. (2015b). Estimating effective landscape distances and movement corridors: comparison of habitat and genetic data. Ecosphere 6(4): 1–16. Mateo-Sánchez, M. C., Gastón, A., Ciudad, C., et al. (2016). Seasonal and temporal changes in species use of the landscape: how do they impact the inferences from multi-scale habitat modeling? Landscape Ecology 31: 1261–1276. Mertzanis, G. (1994). Brown bear in Greece: distribution, present status–ecology of a northern Pindus subpopulation. International Conference on Bear Research and Management 9: 187–197. Mertzanis, G. (1999). Status and management of the brown bear in Greece. In: Servheen, C., Herrero, S. & Peyton, B. (Eds.), Bears (pp. 72–81). Gland: IUCN – The World Conservation Union. Mertzanis, Y. (2000). Incidence of human activities (especially hunting) on brown

270

bear populations in Greece. In: Layna, J. F., Heredia, B., Palomero, G. & Doadrio, I. (Eds.), La conservacion del oso pardo en Europa: un reto de cara al siglo XXI (pp. 117–132). Seville: Fundacion Bioversidad. Mertzanis, G., Isaak, I., Mavridis, A., et al. (2005). Movements, activity patterns and home range of the brown bear (Ursus arctos, L.) in Rodopi mountain range, Greece. Belgian Journal of Zoology 135: 217–221. Mertzanis, G., Kallimanis, A. S., Κanellopoulos, N., et al. (2008). Brown bear (Ursus arctos, L.) habitat use patterns in two regions of northern Pindos. Journal of Natural History 42: 301–315. Mertzanis, G., Korakis, G., Tsiokanos, K. & Aravidis, I. (2009). Expansion of brown bear range in the course of rural abandonment during the 20th century – a case study from the Pindos mountain range. In: Saratsi, E., Burgi, M., Johann, E., Kirby, K. & Moreno, D. W. (Eds.), Woodland cultures in time and space – Tales from the past, messages for the future (pp. 330–337). UK: Embryo Publications. Mertzanis, G., Mazaris, A., Sgardelis, S., et al. (2011). Telemetry as a tool to study dispersal ability, habitat suitability and distribution patterns of brown bears as affected by the newly constructed Egnatia highway – N. Pindos – Greece. In: Krejcar, O. (Ed.), Moden Telemetry (pp. 307–329). London: InTech Open. Mills, L. S. & Allendorf, F. W. (1996). The one-migrant-per-generation rule in conservation and management. Conservation Biology 10: 1509–1518. Morini, P., Pinchera, F. P., Nucci, L. M., et al. (2017). Brown bears in Central Italy: a 15year study on bear occurrence. European Zoological Journal 84: 26–33. Naves, J., Wiegand, T., Revilla, E. & Delibes, M. (2003a). Endangered species constrained by natural and human factors: the case of brown bears in northern Spain. Conservation Biology 17: 1276–1289. Naves, J., Wiegand, T., Revilla, E. & Delibes, M. (2003b). Endangered species constrained by natural and human factors: the case of brown bears in northern Spain. Conservation Biology 17: 1276–1289. Naves, J., Fernández-Gil, A., Rodríguez, C. & Delibes, M. (2006). Brown bear food habits at the border of its range: a longterm study. Journal of Mammalogy 87: 899–908.

Ordiz, A., Bischof, R. & Swenson, J. E. (2013). Saving large carnivores, but losing the apex predator? Biological Conservation 168: 128–133. Ordiz, A., Kindberg, J., Sæbø, S., Swenson, J. E. & Støen, O. G. (2014). Brown bear circadian behavior reveals human environmental encroachment. Biological Conservation 173: 1–9. Ordiz, A., Støen, O. G., Delibes, M. & Swenson, J. E. (2017). Staying cool or staying safe in a human-dominated landscape: which is more relevant for brown bears? Oecologia 185: 191–194. Palomero, G., Fernández, A. & Naves, J. (1993). Demografía del oso pardo en la Cordillera Cantábrica. In: Naves, J. & Palomero, G. (Eds.), El Oso Pardo (Ursus arctos) en España (pp. 55–80). Madrid, Spain. Palomero, G., Ballesteros, F., Nores, C., et al. (2007). Trends in number and distribution of brown bear females with cubs-of-the-year in the Cantabrian Mountains, Spain. Ursus 18: 145–157. Penteriani, V., Delgado, M. D. M., Krofel, M., et al. (2018a). Evolutionary and ecological traps for brown bears Ursus arctos in human-modified landscapes. Mammal Review 48: 180–193. Penteriani, V., Huber, D., Jerina, K., et al. (2018b). Trans-boundary and transregional management of a large carnivore: Managing brown bears across national and regional borders in Europe. In: Hovardas, T. (Ed.), Large carnivore conservation and management: Human dimensions and governance. Oxford: Routledge, Taylor & Francis Group. Penteriani, V., Zarzo-Arias, A., Bombieri, G., et al. (2018c). Density and reproductive characteristics of female brown bears in the Cantabrian Mountains, NW Spain. European Zoological Journal 85: 313–321. Penteriani, V., Zarzo-Arias, A., NovoFernández, A., Bombieri, G. & LopezSanchez, C. (2019). Responses of an endangered brown bear population to climate change based on predictable food resource and shelter alterations. Global Change Biology 25: 1133–1151. Penteriani, V., Zarzo-Arias, A., Delgado, M. M., et al. (2020). Female brown bears use areas with infanticide risk in a spatially confined population. Ursus 2020(31e2): 1–9. Pérez, T., Naves, J., Vázquez, J. F., et al. (2010). Evidence for improved connectivity between Cantabrian brown bear subpopulations. Ursus 21: 104–108.

Human-Modified Landscapes

Pérez, T., Naves, J., Vázquez, J .F., et al. (2014). Estimating the population size of the endangered Cantabrian brown bear through genetic sampling. Wildlife Biology 20: 300–309. PNALM. (2018). Rapporto orso marsicano 2017 Notiziario PNALM-Natura protetta n 21 estate 2018. Available from www .parcoabruzzo.it/pdf/NaturaProtetta_ RapportoOrso2017.pdf [in italian]. Posillico, M., Meriggi, A., Pagnin, E., Lovari, S. & Russo, L. (2004). A habitat model for brown bear conservation and land use planning in the central Apennines. Biological Conservation 118: 141–150. Ripple, W. J. & Beschta, R. L. (2012). Trophic cascades in Yellowstone: the first 15 years after wolf reintroduction. Biological Conservation 145: 205–213. San Miguel, A., Ballesteros, F., Blanco, J. C. & Palomero, G. (Eds.). (2012). Manual de buenas prácticas para la gestión de corredores oseros en la Cordillera Cantábrica. Madrid, Spain: Fundación Oso Pardo (FOA) and Ministerio de Agricultura, Alimentación y Medio Ambiente. Schaal, B. A. & Leverich, W. J. (1996). Molecular variation in isolated plant populations. Plant Species Biology 11: 33–40. Scharf, A. K. & Fernández, N. (2018). Upscaling local-habitat models for largescale conservation: assessing suitable areas for the brown bear comeback in

Europe. Diversity and Distributions 24: 1573–1582. Skrbinšek, T., Jelenčič, M., Waits, L. P., et al. (2012). Using a reference population yardstick to calibrate and compare genetic diversity reported in different studies: an example from the brown bear. Heredity 109: 299–305. Steyaert, S. M. J. G., Endrestøl, A., Hackländer, K., Swenson, J. E. & Zedrosser, A. (2012). The mating system of the brown bear Ursus arctos. Mammal Review 42: 12–34. Swenson, J. E., Gerstl, N., Dahle, B. & Zedrosser, A. (2000). Action plan for the conservation of the brown bear (Ursus arctos) in Europe. Nature and Environment No. 114. Strasbourg: Council of Europe Publishing. Tosoni, E., Boitani, L., Gentile, L., et al. (2017a). Assessment of key reproductive traits in the Apennine brown bear population. Ursus 28: 105–116. Tosoni, E., Boitani, L., Mastrantonio, G., Latini, R. & Ciucci, P. (2017b). Counts of unique females with cubs in the Apennine brown bear population, 2006–2014. Ursus 28: 1–14. Treves, A. & Karanth, K. U. (2003). Human– carnivore conflict and perspectives on carnivore management worldwide. Conservation Biology 17: 1491–1499. Tsaparis, D., Karaiskou, N., Mertzanis, Y. & Triantafyllidis, A. (2014). Non-invasive genetic study and population monitoring

of the brown bear (Ursus arctos) (Mammalia: Ursidae) in Kastoria region – Greece. Journal of Natural History 49: 393–410. Van Gils, H., Westinga, E., Carafa, M., Antonucci, A. & Ciaschetti, G. (2014). Where the bears roam in Majella National Park, Italy. Journal for Nature Conservation 22: 23–34. Vucetich, J. A. & Waite, T. A. (2000). Is one migrant per generation sufficient for the genetic management of fluctuating populations? Animal Conservation 3: 261–266. Wiegand, T., Naves, J., Stephan, T. & Fernandez, A. (1998). Assessing the risk of extinction for the brown bear (Ursus arctos) in the Cordillera Cantabrica, Spain. Ecological Monograph 68: 539–570. Zarzo-Arias, A., Delgado, M., Ordiz, A., et al. (2018). Brown bear behaviour in humanmodified landscapes: the case of the endangered Cantabrian population, NW Spain. Global Change and Conservation 16: e00499. Zarzo-Arias, A., Penteriani, V., Delgado, M. M., et al. (2019). Identifying potential areas of expansion of the endangered brown bear population in the Cantabrian Mountains (Asturias, NW Spain). PLoS ONE 14(1): e0209972. Zedrosser, A., Steyaert, S. M. J. G., Gossow, H. & Swenson, J. E. (2011). Brown bear conservation and the ghost of persecution past. Biological Conservation 144: 2163–2170.

271

Part IV Chapter

20

Conservation and Management

Conservation and Management of Bears Christopher Servheen, Hu¨seyin Ambarlı, Harendra Singh Bargali, Stewart W. Breck, Neil D’Cruze, Claudio Groff, Gabriella M. Fredriksson, Michael L. Gibeau, Issac Goldstein Aizman, Djuro Huber, Katherine C. Kendall, Sterling D. Miller, Michael F. Proctor, Murray Rutherford, Lorraine Scotson, and Jon E. Swenson

Why Bears Need Conservation and Management: Bears as a Conservation-Reliant Species We live in a rapidly changing world where the resiliency of the natural environment is being reduced every day by the actions of humans, whose global population is growing at a worrying rate. A recent analysis reported that land degradation seriously impacts 75% of terrestrial ecosystems and there has been an overall decline of 60% in species population sizes between 1970 and 2014 (WWF 2018). Of 8688 species on the IUCN Red List in the Near Threatened or Threatened categories, 72% are being overexploited for commerce, recreation, or subsistence. Sixty-two percent of these species are also being negatively impacted by agricultural activities including land conversion, conflicts with livestock and crops, and timber plantations (Maxwell et al. 2016). Bears are clearly among those taxa so impacted and threats to bears are increasing. As human activities continue to increase habitat and population pressures on wild bears, we need to recognize that the only way we can continue to maintain bear populations in many areas of the world (especially those with higher human densities) is with continued conservation and management including reducing direct conflicts with humans, maintaining sufficient suitable habitat, controlling mortality to sustainable levels or reducing mortality to allow populations to increase, controlling motorized access to habitat, managing habitat and populations to allow linkage between population units for genetic and demographic rescue, and augmenting or reintroducing populations where necessary. Species requiring such ongoing and continuous management and conservation have been termed “conservation-reliant species” (Scott et al. 2005, 2010). All bear species are certainly conservation-reliant. The healthy North American black bear and most European brown bear populations are healthy throughout most of their range because of sustainable mortality limits established with updated biological data from ongoing research and monitoring. We have a long way to go if we are to establish similar bear monitoring and management systems across the range of all of the conservation-reliant bear species worldwide. The purpose of this chapter is to highlight what conservation and

management can and should be implemented and how for the conservation-reliant bears across the world.

An Overview of the Basis for and Organization of Bear Conservation and Management Schemes Across the World The management of wildlife varies worldwide in both organization and application. Understanding this variation is fundamental to recognizing the strengths and weaknesses in the management of the world’s eight bear species across their range. The relationship of human society to wildlife in general and bears in particular has evolved over time. It may be useful here to use the evolution of North American interactions with wildlife as an example of this evolution of human/wildlife interactions. In much of North America, as well as in some European countries, the management of wildlife including North American black bears, brown bears, and polar bears has evolved through three phases over time: 1. The exploitation phase driven by market hunting for sale of meat and hides; elimination of predators as competitors with livestock; and hunting for subsistence. This phase resulted in the functional extinction of many wildlife species and significant reductions in a wide range of wildlife, particularly predators. The three North American bears declined precipitously under this exploitation phase primarily because of aggressive predator extermination using poisons and unregulated killing. In Europe, the extermination of brown bears was encouraged by government-financed bounties or other predator control efforts (Swenson et al. 1995; Zedrosser et al. 2011). 2. The selective conservation phase for certain species deemed desirable by sport hunters. This was a change in focus from pure exploitation of all wildlife to the maintenance of some desirable species and the habitat they depend upon. This phase saw a change in hunter motivation from subsistence hunting to “trophy hunting” by recreational hunters. This phase of sport-hunter-driven management produced what has been called the “North American Model of Wildlife Conservation” (Geist 1995; Geist et al. 2001; Mahoney 2004). There are ongoing difficulties in the application of

273

Conservation and Management

this model to the conservation of wildlife not thought to be of interest to sport hunters (Organ et al. 2012), and concerns that this model depends on desires for sport hunting by a small segment of the population – mostly male sport hunters (Peterson & Nelson 2016). Some bears such as North American black bears and some brown bear populations benefited during this phase, in contrast to the intensive persecution they suffered during the exploitation phase. This was also the case for brown bears in Sweden, where the population increased rapidly after appropriate conservation measures were introduced, allowing hunting quotas to later increase accordingly (Swenson et al. 2017). Nevertheless, a major difference from the “North American Model of Wildlife Conservation” is that the owner of the hunting rights for most wildlife species in most countries in Europe is the landowner. So, even if a government agency issues the hunting quotas for bears, for example, the hunting-rights owners profit from the hunting (Knott et al. 2014). This results in bear hunting being concentrated into a more restricted group and the economic value of hunting being more important in Europe than in North America. 3. The inclusive conservation phase with enhanced interest in the conservation of all species, not just those of interest to sport hunters. This phase has its origins in the fact that wildlife is of value to the majority of Americans and Canadians, not just hunters. There has been a steady decline in the percentage of Americans who hunt wildlife for sport. Currently only 4.4% of adults aged 16 years and older are sport hunters (US Fish and Wildlife Service 2016). The inclusive conservation phase has been driven by public interest in the existence value of wildlife as opposed to the strictly utilitarian view common to both the exploitation phase and the selective conservation phase. The inclusive conservation phase has seen the rise of Conservation Biology (Soule 1985, 1986) and the passage of wildlife legislation such as the US Endangered Species Act to protect species not associated with hunting. Bears have benefited during the inclusive conservation phase with the recovery of small bear populations, non-lethal management of human–bear conflicts rather than removal of conflict bears, concerns about global warming and polar bears, and the reduction or elimination of sport hunting for bears in some areas. There has been erosion of public support for bear hunting unless it involves consumption of the meat of the bear. In Canada, the province of British Columbia banned grizzly bear hunting in 2017 when 91% of the public opposed trophy hunting of animals such as brown bears. The general public interest in conservation of natural resources has also been increasing in Europe, which is probably an important reason why all of the large carnivore species there are showing an increase over most of the continent (Chapron et al. 2014). Nevertheless, it appears that the poorer European countries have been most successful in conserving brown bears (Kojola et al. 2018).

274

Bear species in most range countries in Asia are fully protected by law, with either no, or very limited, hunting being allowed. Even with this legal protection, in some areas, Asian bear species are in decline and are listed as Threatened on the IUCN Red List. Continued conversion, degradation, and fragmentation of remaining habitat, poaching for body parts, and killing related to human conflicts are key causes of these declines. Although hard data on bear populations are lacking from all Asian range countries, forest cover maps suggest that forest range is declining at a rapid pace or is already at critically low levels, for example in Thailand. The “empty forest syndrome” is an issue in countries like Vietnam and Lao PDR where the few forests that do exist have few native large mammals. Human attitudes toward bears have changed across the world at different rates and levels over time similar to the changes in North America. This does not mean that we have left the exploitation phase behind, as evidenced by the ongoing trade in bear parts in Asia. High prices for gall bladders and lack of or weak law enforcement make Asiatic black bears and sun bears targets for poaching, and in the case of these species it is frequently no longer a matter of tolerance when humans encounter these species outside of protected areas, but more a matter of pure survival skills under high poaching pressures. In some areas of the world bears face continuous conflicts with people due to general public intolerance and crop raiding and livestock depredation. Even in Norway, which is a wealthy country with an advanced system of wildlife management and a generous compensation program for bear-caused depredations on sheep, the conflict level is very high (Gangaas et al. 2013). In fact, a generous compensation system can be counterproductive and actually reduce the efforts of livestock owners to prevent livestock losses. Worldwide we are entering the inclusive conservation phase for many bear populations with increasing interest in managing the issues that cause human–bear conflicts rather than just eliminating bears. There is a big difference between documenting the problems for bear conservation and management and successfully solving these problems with effective actions. There has been progress in identifying the root causes of bear population declines for many species from loss of sea ice for polar bears to excessive exploitation for trade in bear parts, to expanding human populations that settle in bear habitat. However, much remains to do to address the causes of these problems and improve the status of bears with effective conservation and management programs in most areas of the world. The successful conservation and management of bears is dependent on effectively addressing four factors (Figure 20.1). All too often in conservation programs, there is a tendency to focus on the collection of biological data as the primary way to conserve a species. However, without attention to developing an organized and funded management team and nurturing both public and political support, bear conservation programs are doomed to fail. In North America and in many areas of Europe, progress has been made in building all four of these

Conservation and Management

Figure 20.1 The four factors that contribute to successful bear management and conservation. Each must be addressed if any management and/or conservation program is to be successful (adapted from Servheen 1998).

factors to construct successful bear conservation and management programs. However, there continue to be setbacks in the United States and in Europe when political support wavered or changed, such as in the lack of political support to address climate change and its ultimate impact on polar bear conservation. In Europe the general situation for brown bears is good, as mentioned above; however, there are still many small and isolated populations that will require intensive management for many years to come, if they are to survive (Chapron et al. 2014). Yet, the level of conflict is high in many countries, and some have adopted a regional management approach, to allow decisions regarding bear hunting and management to be made at a more local level, within national goals, in order to reduce the level of conflict (Redpath et al. 2017; Figure 20.2). The development of public and political support for bears is particularly a problem in much of southern Asia where effective bear conservation and management is in its infancy, with many charismatic species facing extinction (e.g. tigers, rhinoceros, and orangutans) and range country governments failing or unable to take timely and appropriate conservation action. There are few South Asian government management agencies with knowledge of or interest in bear conservation and very few skilled bear managers to address human–bear conflicts and build public support and understanding for bear conservation. Building and maintaining political and public support for bear conservation requires continuous effort because the political interests and the perceptions of people can and do change. Political and public support are closely related because political support is generally determined by public support or by pressure from outside groups. Progress on South Asian and Andean bear (Tremarctos ornatus) conservation and management faces challenges in addressing these four factors. Bear management is not present in terms of American-style wildlife management in the Middle East because the brown bear is a protected species in most countries and Wildlife Departments try to protect the species against poaching. There are some efforts to offer either some mitigation measures such as implementation of electric fences or wildlife damage insurance systems. An example is that half

Figure 20.2 Brown bear in British Columbia, Canada, being hazed with bean bags, cracker shells, and rubber bullets upon release after being relocated from a site where the bear obtained food at a human residence. Hazing is an effort to aversely condition the bear from human presence in the future. (Photo by Dave Quinn.)

the cost of premium insurance is subsidized by the Ministry of Agriculture and Forestry in Turkey (Tarsim 2018). If resentment among some local people arises after severe bear-caused damage, the Turkish wildlife department can adjust the hunting quota for provinces suffering this damage. However, such hunts are rarely effective in reducing conflicts because hunters are mostly foreigners who target large bears and not necessarily the individuals causing damage. In summary, we have a long way to go to build effective conservation and management programs for bears in many areas of the world. Overcoming these challenges will require cultural sensitivity and an organized effort in both developed and developing countries if we are to secure a future for the bears of the world.

Habitat Conservation and Management Habitat Management Habitat quality and security from human-caused mortality are two of the most important aspects of bear conservation management in North America and likely around the world. They bring together the two driving forces thought to control all wildlife populations: bottom-up and top-down population influences. Habitat quality relates to the necessary food and energy resources important for reproduction and successful hibernation in boreal species. Delayed implantation and hyperphagia are evolutionary adaptations designed to allow bears to survive as a species through undulant interannual food supplies. There are often several natural food resources that are disproportionately important for bears’ capacity to store fat resources to fuel hibernation, successful reproduction (Robbins et al. 2012; Hertel et al. 2016), and survival through low fruit productivity years in the tropics (salmon, ungulate calves, berries, acorns, figs, termites and more; Hilderbrand et al. 1999; Hashimoto et al. 2003; Schwartz et al. 2014;

275

Conservation and Management

Figure 20.3 Brown bear at a feeding site in Croatia. The barrel contains corn and has holes in it to minimize the amount of food obtained by the bear, thus prolonging the bear’s presence at the shooting site. Hunters sit in the shooting tower and shoot bears that come to feed at the site. (Photo by Mihajlo Kovačević.)

McLellan 2015; Hertel et al. 2018a). Seasonal food resources are, of course, important, but late summer and early fall (or productive seasons in more tropical areas) hyperphagia foods are the determinants of northern bear reproductive success (Schwartz et al. 2006; McLellan 2011, 2015; Proctor et al. 2017; Hertel et al. 2018a). For bears living in more subtropical regions, where food is available year-round, fluctuations in the scale of fruiting events can play a large role in survival and life histories (Wong et al. 2005; Fredriksson et al. 2006). Most species of bear are very susceptible to excessive human caused mortality (McLellan et al. 1999; Nielsen et al. 2004; Steyaert et al. 2016; Penteriani et al. 2018), as they generally have low reproductive rates that include late maturation, long interbirth intervals, and small litter sizes (Miller 1990; Schwartz et al. 2003; Garshelis 2009). Bears seek out easily digestible foods such as meat, fish, insects, berries, and other shrub fruits, and tree fruits (Pritchard & Robbins 1990; McLellan 2011). These needs drive bears toward human-based foods such as agricultural crops and livestock that result in human–bear negative interactions and in many cases excessive human-caused mortality (Figure 20.3). The top-down influence on bears is usually human-caused mortality (McLellan et al. 1999; Nielsen et al. 2004; Steyaert et al. 2016). This is directly related to the probability of interaction between bears and humans: either bears seeking foods in human-dominated landscapes, or humans seeking foods or recreation in bear habitats. If it is sun bears (Helarctos malayanus) being drawn to oil palm plantations (Nomura et al. 2004; Fredriksson 2005), Andean bears drawn to maize (Can et al. 2014), or polar bears (Ursus maritimus) attracted to whale carcasses near northern seaside villages (Derocher et al. 2013; Atwood et al. 2016), the consumption of these energy-rich, easily acquired foods usually results in significant mortality risk. This mortality risk is aggravated by the fact that bear

276

habitats have been lost across the world as humans settled and exploited lands that were previously natural habitats. Human presence in these areas converts them from natural food production to concentrated abundances of human-related energyrich food for bears such as corn (maize), palm nuts, grains, and fruit trees. Depending on public attitudes and local norms toward human–bear coexistence, natural habitats can thus become mortality sinks for bears (Penteriani et al. 2018). While managing front-country, human-dominated landscapes for reduced conflict mortality, it is just as important to manage habitats away from humans for reduced mortality risk, especially as connectivity of bear populations needs to be maintained over the overall landscape. This means that to be effective, mortality management must include both humandominated landscapes and natural habitats. In South-East Asia, where wildlife law enforcement in most countries is low at best (Müller & Kaji 2016), bears are killed and poached for body parts, specifically gall bladders which reach high values at Traditional Chinese Medicine (TCM) markets (Foley et al. 2011). With this high poaching value and low law enforcement, bears are specifically targeted within protected areas, and bears are neither safe nor tolerated when venturing outside these protected areas.

Habitat Quality Providing for habitat quality is very important but can often be challenging. Managing for habitat quality may take many forms, from the creation of a protected area network, which is more than just unproductive but beautiful landscapes (rock and ice in North America have traditionally been strong candidates for parks; Joppa & Pfaff 2009). Other resources may require management across a spectrum of habitats, such as healthy salmon stocks as bear food requires salmon management in spawning streams, but also in ocean environments and migration routes. This is the case for much of the Pacific basin from Hokkaido, Japan, to Kamchatka, Russia, north-east across the north Pacific to the famous coastal fish-bearing streams, and the rivers of the Alaska and British Columbia coasts. US west coast states of Washington, Oregon, and California currently do not support a healthy salmon–bear ecosystem due to decades of excessive mortality where bears have been extirpated regionally (Mattson & Merrill 2002). This is an example of both the habitat quality and security not being looked after. So not only do those areas need healthy salmon-bearing streams and rivers, there must also be some level of limited human access to a portion of these resources so bears may use and benefit from these resources without undue mortality risk. Managing for other high-quality hyperphagia foods may come in the form of long-term land-use management. For example, important berry patches (Vaccinium spp.) in portions of western North America become more productive for bear foods after wildfire events (McLellan & Hovey 2001a), therefore managing for long-term berry supplies for bears may entail shifts in fire suppression patterns designed to save forests for timber harvests (McLellan & Hovey 2001a).

Conservation and Management

High-productivity forests in Asia are usually in the lowlands and near rivers, areas where human populations are highest, and as a result, few of these important areas remain as wildlife habitat. Managing high-quality habitats for tropical bears in Asia is now more related to establishing and managing protected areas, in remnant forests (e.g. usually in rugged mountains), reducing excessive forest harvest, halting conversion of natural forest to palm oil plantations, forest fire prevention, and retaining or restoring connectivity of landscapes between protected areas (Wong et al. 2013; Scotson et al. 2017b). Another aspect of habitat management is managing for sea ice so polar bears can access rich seal hunting grounds. This is a global issue related to reducing worldwide CO2 production to minimize climate change and we might be losing that battle (Derocher et al. 2013).

Security from Human-Caused Mortality Security from human-caused mortality is well developed and documented in western North America for the grizzly bear (Ursus arctos horribilis). Here we discuss the principles that have been shown to result in successful habitat management and conservation successes. The southern extent of the North American brown (grizzly) bear distribution in the north-west lower 48 United States, and southern British Columbia and Alberta, Canada, has been a contentious arena for good habitat management for the past 40 years. After the conservation paradigm shift that occurred in the late 1970s, habitat management has been applied successfully in portions of this subcontinental region, with various degrees of intensity and success. A closer look reveals valuable lessons. Grizzly bears of the North American interior have relatively large home ranges that can bring them into contact with people. Their occasional aggressive nature exacerbates their susceptibility to excessive human-caused mortality by stimulating a quicker lethal response from people (Mattson & Merrill 2002). These traits result in significantly higher mortality risk in habitats with higher back-country road densities. In North America, it has been shown across many types of habitats that higher forest road densities are associated with lower female bear survival rates (Mace et al. 1996; Schwartz et al. 2010; Boulanger & Stenhouse 2014; Steyaert et al. 2016; Proctor et al. 2017; Lamb et al. 2018). Limiting motorized access to high-quality habitats has been a cornerstone of recovery management for threatened grizzly bear populations in North America for decades and has generally been successful (Kendall et al. 2009; Schwartz et al. 2010; Mace et al. 2012). These areas have relatively high percentages of protected areas within their ecosystem borders and managers had the luxury of starting conservation efforts with hundreds of bears. A major part of the conservation program for these populations was a comprehensive program of motorized access management to close thousands of kilometers of existing timber harvest roads and improve habitat security. Motorized access management has also had a beneficial effect on the status of the

isolated and much smaller ( 1.14) yet documented for the species (Sæther et al. 1998) and has a relatively high population density of 30 bears per 1000 km2 (Bellemain et al. 2005). In this system, berries, especially bilberries (Vaccinium myrtillus), are the most important food during hyperphagia (Stenset et al. 2016). Even though the number of bilberries in a given year affects the growth and reproduction of the bears (Hertel et al. 2018a), bears do not come closer to human settlements during years of poor berry production (Hertel et al. 2018b). In fact, bears in Sweden rarely approach human settlements to obtain human-derived food (Elfström et al. 2014). This contrast in the response of brown bears to habitat management differences between North America and Sweden highlights the importance

277

Conservation and Management

of recognizing the diversity of response that bears (even within the same species) can exhibit to human impacts across their range. It also emphasizes the need to avoid blanket application of types of habitat management as a solution to bear conservation. We must recognize that habitat management is dynamic, involving many factors including the response to and tolerance of humans toward bears, the adaptability of the local bear population, and factors such as habitat productivity that will influence how bears can survive across the world in response to human actions. In Latin America the landscape-scale conservation approach for the conservation of the Andean bear has been implemented in Colombia, Ecuador, and Peru. In Colombia, The National Natural Parks System (PNN & WCS 2018) has implemented an Andean bear conservation strategy based on the conservation of wilderness areas and the reduction of human–bear negative interaction. The goal of the strategy is the conservation of long-term viable population by managing and conserving large landscapes composed of interconnected national parks, wilderness areas, and human dominated habitats. In Ecuador and Peru (SERNANP et al. 2014), the Andean bear conservation strategy focuses on a protected area mosaic approach where human productive activities both outside and within protected areas are managed to lower their vulnerability to bear activity and reduce human–bear negative interactions. While there are caveats, which include regional habitat quality, cultural differences, tolerance and lethality of humans, and political differences, several general principles of habitat management are applicable across systems and species: • Humans and bears can coexist as long as human activities are managed to limit attractants that bring bears into conflict with people resulting in excessive mortality risk and there are sufficient natural habitats protected from human settlement and agriculture. Habitats do not need to be devoid of humans to be of value to bears. Human activity must be managed in high-quality bear habitat where bears seek high-energy natural foods, usually on a seasonal basis. • Protected areas where human landscape disturbance and human-caused mortality risk are both low are an important aspect of a successful habitat management system for most bears. Protected areas are especially important in tropical habitats where human tolerance for bears is low. In tropical bear habitats, there must be representative amounts of protected habitat containing a diversity of tropical forest habitats to serve as reserves where bears can exist with little mortality risk due to poaching and human presence. • Understanding the variability of habitat productivity for important bear foods is very useful to inform and guide where and how to manage natural habitats. It is particularly useful to understand where the most important hyperphagia foods are located, so those areas can be managed (in some cases by seasonal road closures) to

278





minimize risk of excessive mortality risk or habitat displacement of bears by human activity. Understanding non-hyperphagic seasonal habitats of importance that bring humans and bears together can be important for managing mortality risk. For example, these may be areas of spring range where bears go after exiting their dens and are concentrated in low elevation, snow-free habitats. Motorized access management is important to maintain habitat security to reduce negative human–bear interactions, at least in North America. One way to do this is to manage densities of roads. One well-documented example of motorized access management exists in western North America where road densities are managed to assure that 50–60% of natural habitat is more than 500 m from open motorized roads and, where roads occur, road density does not exceed 0.5–0.6 km/km2. This may not apply in some European areas such as in Sweden where high road densities seem to have little impact on brown bear population health. Another example of motorized access management is to close timber harvest roads when timber harvest operations are complete. In much of North America, proper forest management for bears requires both (1) management of the location and type of logging to provide diverse natural foods; and (2) closure of timber harvest roads upon completion of timber harvest to maintain habitat security.

The Role of Protected Areas in Bear Conservation Protected areas now cover between 12% and 15% of the Earth’s terrestrial surface depending upon how broadly one defines “protected area.” The common purpose of these protected areas is to sustain ecosystems and all their constituent species as well as to maintain ecosystem services and functions. Another way to think about the role of protected areas is to avoid, prevent, and mitigate against anthropogenic conversion of natural habitat. It is this confluence of maintaining natural habitats and avoidance of anthropogenic change that makes protected areas an important element in the conservation of the world’s bears. The conversion of natural habitats into human-dominated land use is considered to be the biggest threat to world biodiversity (Pereira et al. 2010). To counteract this phenomenon, the protected area concept has become a prominent conservation strategy in the last century. This designation of protected areas arguably constitutes “one of the most stunning conservation successes of the twentieth century” (Ervin 2003). While the concept of protecting natural habitats is a primary purpose depending upon jurisdiction, the management reality can be significantly different on the ground. Globally, relatively few protected areas are managed with biodiversity conservation as the primary objective. Of the world’s 98,400 terrestrial

Conservation and Management

protected areas, only 8800 (8.9%) are listed under IUCN categories I or II (Naughton-Treves et al. 2005). Indeed, many protected areas were not originally chosen for their species abundance or richness. Caughley (1994) suggested that species end up not in the habitat most favorable to them, but in the habitat least favorable to the agent of decline. Concurrent with the rise of establishing new protected areas in the last quarter century, and especially in developing countries, the mission of protected areas has expanded from biodiversity conservation to improving human welfare (Naughton-Treves et al. 2005). This dual mandate poses challenges for the ecological value of these areas for bears, and all other wildlife for that matter. Woodroffe and Ginsberg (1998), among others, point out that even when habitat is otherwise suitable, in isolation many protected areas simply are not large enough to maintain viable populations of species such as bears that occur at low densities. In order to maintain populations of large mammals, protected areas need to become larger if they are to be fragmented from similar habitat in the surrounding landscape. More importantly, attempts to link protected areas need to be embedded within a broader concept of landscape conservation, in which protected areas are but one component, and in which protected areas are integrated into larger landscape planning and management frameworks (Gaston et al. 2008). In theory, a number of models address this broader view of landscape conservation including: (1) community-based conservation, where partnerships link biodiversity objectives with local development objectives; (2) conservation easements, where private landowners voluntarily restrict the types of activities on their property to enhance conservation values; and (3) comanagement, where government shares power and decision making with local authorities to reach common goals. In many areas of the world, however, such as many countries in Asia, the fate of bears rests increasingly on their persistence in protected areas. Outside of these protected areas, bears are subject to accelerating habitat loss due to conversion to plantations, degradation of habitat due to overharvesting of timber, human encroachment, and specifically human intolerance and direct persecution for their body parts. Many of these threats also occur inside protected areas due to weak laws or absence of law enforcement. The future of bears in many developing countries, particularly in much of Asia, is increasingly uncertain because effective protected areas are few, the most productive forest areas are converted to other purposes, and now forest mostly persists in mountainous regions, which may not be the optimal habitat for some of the bear species. It is apparent now in the twenty-first century that the blueprint approach of protected areas is overly simplistic to provide the only solution to species conservation in an increasingly complex world. Likewise, the panacea of communitybased conservation is probably no more effective than the panacea of exclusively state-based conservation (Berkes 2007). Conservation solutions are increasingly elusive in a globalized world. A more strategic approach (Servheen 1998)

is warranted if we are to conserve the different bear species we currently have around the globe.

Population Fragmentation and its Management: The Importance of Population Connectivity to Bear Conservation Population fragmentation is ubiquitous in modern landscapes and is a threat to biodiversity in general (Wilcove et al. 1998; Fahrig 2003; Wilson et al. 2014), and specifically to bears (McLellan et al. 2017). Extirpation risk due to the small population paradigm of isolated subpopulations (Caughley 1994) can be a mechanism of range contraction and was a likely contributor in the case for brown bears across North America in the ninetenth century (Mattson & Merrill 2002). Beyond these historic and current conservation threats, the emergence of climate change elevates the importance of connectivity conservation substantially. In a review of 22 years of climate change mitigation tools, reversing fragmentation through connectivity management was the most recommended strategy (Heller & Zaveleta 2009). We need to allow species and ecosystem components the ability to move in response to shifting habitats to survive the significant yet unpredictable changes related to climate change (Penteriani et al. 2019). There are two primary types of fragmentation and its antithesis, connectivity: genetic and demographic. Genetic fragmentation results in a loss of genetic diversity through increased inbreeding, which occurs faster in small populations, and potentially threatens a species’ ability to adapt to relatively rapid environmental changes (Keyghobadi 2007). While traditionally this had been a longer-term process (many generations), the rapid environmental changes we are experiencing, and are expected to experience, from climate change may hasten the effects of genetic fragmentation. Demographic fragmentation, the interruption of movements of each, or either, gender and the impacts that result, can be a more urgent concern. For example, an isolated small population with no immigration from either sex has no ability for demographic rescue that can result from female immigration (Proctor et al. 2005, 2012). Stochastic (chance) mortality patterns may result in excessive female mortality, leaving a population without the ability to grow fast enough to survive. Demographic rescue, in the form of female immigrants, would offer the potential to elevate reproductive rates for population increase (Proctor et al. 2005, 2012). The example of brown bears in North America illustrates these scenarios. Because male-biased dispersal in bears often mediates gene flow (McLellan & Hovey 2001b; Proctor et al. 2004; Zedrosser et al. 2011), traditional measures of genetic discontinuity will usually miss fragmentation that primarily affects females but not males (Proctor et al. 2005). Natal dispersal in brown bears in North America is male-biased (McLellan & Hovey 2001b; Proctor et al. 2004), and therefore assessments of genetic discontinuity often cannot detect female

279

Conservation and Management

fragmentation in systems where female movements are limited but not males’. This is the case for brown bears (Proctor et al. 2005, 2012) and black bears (Proctor et al. 2020a) in North America. However, using individual-based genetic analyses that follow the fates of individuals has allowed the detection of female fragmentation even when there is no sign of genetic discontinuity (Proctor et al. 2005, 2012, 2020a). These situations are a result of the fact that human activities more easily restrict female brown bear movements as they naturally have shorter natal dispersal. The short natal dispersal distances of female brown bears, where one generation’s new home range overlaps some portion of their mother’s home range (McLellan & Hovey 2001b), means that for female (demographic) connectivity to occur, it usually takes several generations. Female bears must therefore live in the connectivity habitats linking two populations, at least, which often entails overlapping with human settlements and transportation corridors that were fragmenting the population in the first place. These are the areas where long-term patterns of human-caused mortality caused the fragmentation. Several efforts within North America have predicted brown bear linkage habitats using habitat corridor modeling from GPS telemetry to inform landscape connectivity management (Proctor et al. 2012; Peck et al. 2017). Connectivity management often takes the form of actions that reduce mortality risk of bears that are spending considerable time sharing habitats with human settlements and the related disturbance (Proctor et al. 2018b). Brown bears in Europe are almost irreversibly fragmented into 10 separate populations. Except in the northern parts of the continent the extant populations are mostly confined to the mountain ranges. With the ever-growing highway and rail enhancements, and other infrastructure like wind turbines and ski resorts, the fragmentation of European bears is continuing. The European Union is pushing toward populationlevel management (Linnell et al. 2008; John et al. 2016) and efforts are underway to attempt to make new and existing infrastructure permeable to bears and other wildlife. Since 1999, in Croatia over 1000 km of new highways has been constructed in bear range. The environmental impact studies for these projects required construction of numerous overand underpasses including 13 green bridges 100–200 m wide (Kusak et al. 2009). The good news is that in spite of all odds, brown bears and other large carnivore populations in Europe have been increasing over the past 20 years, indicating that with wildlife-friendly legislation and careful management, brown bears and other carnivores can survive and even increase in the heavily settled landscape of Europe (Chapron et al. 2014). Another example comes from polar bears. The potential fragmentation of polar bear populations has been thoroughly assessed and the results show that they are experiencing very little human-caused fragmentation. Subpopulation boundaries have been delineated with telemetry and genetic methods (Paetkau et al. 1999; Malenfant et al. 2016). Discontinuities

280

are most often associated with large landmasses and to a lesser extent large ice-free oceanic expanse where inter-area movements are relatively low. Climate change has been ushering in a period of declining sea ice, which has resulted in decreases in polar bear body condition (Obbard et al. 2006), and potentially changes in gene flow patterns in portions of their range (Peacock et al. 2015) but not yet extensive population declines or habitat loss. Because anthropogenic forces do not cause fragmentation of polar bears into their 19 circumpolar subpopulations (see Chapter 14), connectivity management is not thought to be part of the response to climate change as is likely the case for other bear species. Sloth bears (Melursus ursinus) are endemic to the Indian subcontinent, and inhabit mostly dry forests from Nepal, through India, into Sri Lanka. Human settlements, rural farming areas, transportation corridors, and cities and towns extensively fragment their populations across India and its dense population of over 1.2 billon people (Garshelis et al. 1999a; Bargali et al. 2012). Human–bear conflict is extensive (Sharma et al. 2013; Ratnayeke et al. 2014; Dutta et al. 2015), often resulting in human injury (Bargali et al. 2005; Singh et al. 2018). There are two levels of challenges that fragmentation and connectivity conservation face in developing countries. First, is that the low level of scientific effort to address fragmentation is due to funding constraints. Radio-collaring is uncommon for sloth bear studies in India, but its use is growing (Ratneyeke et al. 2007; Bargali et al. 2012). Genetic studies that reveal patterns in fragmentation are being initiated (Dutta et al. 2015). The second challenge is dense human populations and expansion of rural farming and human settlement areas into bear habitat that fragments wildlife habitats and populations. The extent of sloth bear fragmentation is increasing (Dutta et al. 2015). While protected areas provide some degree of security and may be the backbone of what might be termed a sloth bear metapopulation, many sloth bears live outside protected areas (Garshelis et al. 1999b). Maintaining movement opportunities for these bears has the potential to secure functional connectivity between protected areas (Dutta et al. 2015). There is some evidence of occasional sloth bear movement between larger habitat patches that are separated by up to 300 km (Dutta et al. 2015). Enhancing connectivity for sloth bears in India requires identification and management of functional corridors between protected areas to allow sloth bears to live in them. The time available to address connectivity for sloth bears is growing short because of climate change and increasing human populations and development in the Indian subcontinent. Giant pandas are another example of the importance of fragmentation and connectivity research and management. Panda fragmentation has been studied relatively extensively compared to other bear species. They have a relatively limited diet (bamboo species) and range and are more susceptible to climate change habitat shifts (Gong et al. 2017) than any other bear species. Panda conservation effort has been significant

Conservation and Management

and has greatly benefitted the species (Loucks et al. 2001; Liu et al. 2016). In 2016, their Red List rank went from Endangered to Vulnerable (Swaisgood et al. 2016), but new threats are emerging, including the expansion of livestock grazing leading to degraded habitats, infrastructure development, and increasing habitat fragmentation discussed further below (Swaisgood et al. 2018). Although fragmentation is recognized as a serious issue to panda conservation, efforts to manage for connectivity between fragmented populations are just beginning. Efforts are underway to identify core habitats, establish corridors, and prioritize areas for protection and restoration. Habitat-use data and habitat-quality models have informed the creation

Figure 20.4 A brown bear going under a barbed wire fence around a bait site. The bear passively leaves hair on the barbs and this hair can be used for analysis of DNA to identify the bear. Such hair collection fences are used for population size and trend analysis, relatedness analysis, to assess connectivity between populations, and other valuable research. (Photo by Stefan Himmer.)

of 67 protected areas (Swaisgood et al. 2018). Genetic analyses have been used to estimate population size, determine population structure and fragmentation patterns, and genetic diversity, understand dispersal patterns and then predict connectivity habitat patterns (Wei et al. 2015; Figure 20.4) and the efficacy of interpopulation corridors (Wang et al. 2014; Wei et al. 2015; Swaisgood et al. 2018). Captive breeding and assisting gene flow by moving animals between fragments has been evaluated to deliver demographic rescue to isolated small populations. Despite all this conservation effort, threats to pandas continue, with expansion of livestock grazing and infrastructure development that erode panda habitat (Wang et al. 2015; Liu et al. 2016; Swaisgood et al. 2018). Pandas might be an example where the maintenance of genetic diversity to preserve the ability to adapt to climate change may be very real and as important as demographic connectivity (Gong et al. 2017). Bamboo habitats are predicted to shift and providing suitable corridors to shifting habitats will be important. China’s successful captive breeding program of pandas may also have a role to play in reintroductions into areas in need of genetic rescue and population supplementation. In summary, our progress in managing connectivity across bear species varies tremendously. The four key drivers of anthropogenic fragmentation are human-related mortality, habitat loss, road network development, and habitat degradation (Figure 20.5). All will intensify in the future with increasing human numbers combined with climate change. These trends can only be reversed through understanding and addressing the spatially explicit causes of fragmentation. It is Figure 20.5 Schematic of mechanisms of bear response to roads. The main effect of mortality ultimately reduces density. Secondarily, habitat displacement and direct habitat loss potentially affect reproductive output and density. (Adapted from Proctor et al. 2020b).

281

Conservation and Management

important to realize that relying on protected areas alone to provide habitat and connectivity for bears, while necessary, is not sufficient in itself to produce conservation solutions in an increasingly globalized world. Connectivity management must be implemented in the matrix of protected areas and the human-impacted habitats that surround protected areas if it is to be successful. These selected examples in this section cover the spectrum of variability of the state of fragmentation understanding and connectivity management for bears. A more thorough treatment of this topic can be found in Chapter 22, which reviews connectivity for all bear species.

Working with People to Develop Tolerance and Support for Bears and Bear Conservation No matter how scientifically sound and well-intentioned a conservation program may be, it is unlikely to be successful if it does not have the support of the people whose behavior, property, or other interests will be affected. Support for bear conservation depends on people’s values, beliefs, and attitudes about bears. It also depends on their attitudes about the agencies or other institutions behind the program, expectations about how the program will affect their broader values, and feelings about the decision-making processes through which the program is developed and implemented. Policy scientists call these factors the “social process” and the “decision process” – the sociopolitical context for a program or policy and the governance structure under which it is designed, approved, and instituted (Lasswell 1971; Clark 2002). People’s values, beliefs, and attitudes are key factors in the social process that underlie their support for or opposition to bear conservation. Research shows that people’s values for wildlife are complex and multidimensional, and that these values vary with demographic factors such as age, education, and occupation (Kellert 1996). Several typologies have been developed to classify values of wildlife and nature (see, for example, Rolston 1988; Kellert 1996; Jones et al. 2016; Manfredo et al. 2017). These typologies distinguish dimensions of value such as the extent to which people believe that wildlife should be used, managed, or dominated, and the extent to which people believe that wildlife have rights and deserve respect, caring, and trust (Manfredo et al. 2017). In addition to the values that people hold for wildlife and nature, their attitudes toward bears are also influenced by factors such as perceptions of the particular species or subspecies involved, knowledge of wildlife biology, and historical and current relationships with bears (Kellert 1994). The latter may include cultural traditions and practices, accepted patterns of utilization such as hunting, history of conflicts with people and property, and conservation status. Of course, these factors will vary with the setting in which a bear conservation program takes place. Although many people have an affinity toward bears, and research in North America and Europe shows a reasonably high degree of support for bear conservation, some

282

people do have negative attitudes about bears and bear conservation, especially in non-urban agricultural areas where dominionistic and utilitarian values may be prevalent and there may be a history of livestock depredation or safety concerns (Kellert 1994; Dressel et al. 2015). Specific attitudes about bears are just one of the relevant components of human perspectives in bear conservation. Conservation initiatives may also affect many of the broader values that people hold and pursue in their lives. Lasswell (1971) identifies eight main types of value at stake in human interactions: wealth, power, respect, well-being, enlightenment, rectitude, affection, and skill. The following brief hypothetical example illustrates how this wider array of values may come into play in conservation and interact with specific attitudes about bears. A decision to prohibit bear hunting in a region may be supported by conservationists because they hold moralistic and humanistic values toward bears, and they perceive the hunting prohibition as aligning with their values of rectitude (doing the right thing) and enlightenment (in their view the decision is supported by the scientific evidence). The same hunting prohibition may be opposed by guide outfitters, because their dominionistic and utilitarian values support the belief that it is not ethically or morally wrong to hunt bears (rectitude), and they predict that the prohibition will cause them to lose guiding fees (wealth) and opportunities to use their professional expertise (skill). Meanwhile, livestock producers such as ranchers may also oppose the prohibition because they believe that the absence of hunting will lead to increased property damage and depredation on livestock (wealth, respect, affection) and more safety risks for their families (well-being, affection). They may also believe that their lifestyles and livelihoods are being disrespected (respect). All of these participants will feel either indulgences or deprivations of power and respect, depending on whether they perceive that their views have been heard and they have been able to influence the decision about hunting. And all of these values and beliefs of individuals are likely to be reinforced by the social roles and groups with which they identify (for example, as livestock producers, conservationists, wildlife scientists, guide-outfitters, or hunters). These complex value-driven dynamics of the social process and decision process are further complicated by the symbolic significance of bears. Bears are potent symbols of wilderness, power, strength, and awe. However, bears and bear conservation also often become symbols in political struggles about deeper issues and changes in society. Conservation programs may provoke anxieties and become the focus in disputes about economic loss, changing lifestyles and landscapes, recreational access, appropriate use of public lands, individual rights versus collective rights, local interests versus national interests, and other highly contested and emotional issues. People use bear conservation programs as an outlet for their frustrations with top-down governance and edicts from afar, agencies that they perceive as not listening, and decisions that they consider unfair and disempowering. If public decision-making

Conservation and Management

processes (governance) do not provide opportunities for people who are affected by conservation programs to be heard and to work through such conflicts to find common ground, their disputes will likely become more bitter and their positions more polarized and entrenched. Conflict can escalate to the point that participants do not see realistic options for finding common ground, and instead focus on gaining more power to advance their own interests (Rutherford & Clark 2005). In such “politicized” settings, wildlife scientists may find that people are not really interested in improving their knowledge and understanding. Rather, people may use scientific research and recommendations only to the extent that they support their personal interests and objectives. Otherwise, they may attack not only the scientific methodology and results, but also the qualifications and integrity of the scientists (Gibeau 2012). Given this highly complex, contested, and potentially hostile sociopolitical context in which wildlife scientists and managers must operate, how can they work with people to increase tolerance for bears, and develop conservation programs that are more broadly supported? The first step is for wildlife professionals to be self-reflective and to clarify their own standpoints – the values, cognitive biases, conceptual frames, and other factors that shape how they see the world, how they conduct their research and how they interact with others (Clark & Wallace 1999). By clarifying their own standpoints, they may be better equipped to understand the perspectives of others, and to see how and why their own views and recommendations may differ from those of the people who are affected. A second recommendation is that wildlife professionals should map the sociopolitical context and decision-making processes that exist in the setting for a conservation program. As the discussion above makes clear, context matters, and bear conservation programs should be designed and adapted to fit the specific sociopolitical settings in which they are instituted. Local leaders, culture, institutions, traditions, decision-making practices, and other factors are all important. Third, wildlife professionals should work with local people to address the real problems they face arising from bears and bear conservation programs. By working on tractable smaller problems, it is possible to build relationships and trust, develop capacity and begin to depoliticize situations, while implementing practical measures to protect bears and reduce conflicts (Wilson et al. 2014). Examples include installing deterrence measures such as electric fencing for beehives and calving sites, helping with changes in agricultural and husbandry practices such as cleaning up bone yards and protecting feed storage facilities, and installing bear-proof food caches or food poles in the back country and garbage facilities in the urban–wildland interface. Fourth, people are more likely to support conservation programs if they are meaningfully engaged in the development of those programs. When wildlife professionals and their agencies are willing to share power and participate in

collaborative decision-making with people who are affected or interested, there is a real possibility to develop robust and effective conservation programs (Gibeau 2012). Essentially, this type of collaborative decision-making provides a setting for dialogue, in which people are encouraged to reason about problems and discuss strategies to address them, and jointly identify the common interest (McLaughlin et al. 2005). Interest-based discussion rather than positional bargaining is encouraged. An initial matter on which people may find it easy to agree is the standards for good decision-making. Even when people initially disagree on what should be done, they may be able to agree on how decisions should be made (Rutherford et al. 2009). As trust improves, scientific research can be reviewed, and the implications discussed in a less politicized environment (Gibeau 2012). Such collaborative decision-making is recommended for all settings, but it is essential for settings in which indigenous people are involved. As Housty et al. (2014, p. 70) observe about the grizzly bear monitoring program they developed in the traditional territory of the Heiltsuk First Nation in Canada – a collaboration that included academics, indigenous government and community leaders, provincial and federal governments, and others – “Successful resource management by Indigenous people may require approaches like this one, in which collaborative science-based management is embedded within a socially and culturally appropriate framework for action.” People are at the center of all conservation efforts. By (1) being more self-reflective, (2) understanding the social context and decision-making processes, (3) working with local people, and (4) engaging people in meaningful dialogue, wildlife professionals can set the foundations for comprehensive problemsolving. Then, by systematically evaluating and learning from trials of such innovative approaches to conservation, they can harvest the lessons and communicate them to others, encouraging replication or adaptation in other settings.

Management of Bear Mortality Mortality documentation is extremely important for management of the world’s bear species. The numbers and ranges of all species have been significantly reduced, and excessive mortality, which is ongoing for all bear species, has been an important cause, except in American black bears, polar bears, and, in many areas, brown bears. Compared to other large mammals, high mortality rates for bears are problematic because bears generally have lower reproductive rates. Unsustainable mortality rates in bears most commonly result from killing bears that cause damage to human interests, such as crops or livestock, of bears perceived to be threats to humans, in accidents, or for commercial sales of parts (termed “non-sport” kills in contrast to legal hunter kills). Non-sport kills may also be numerically larger than hunting kills in some areas and may also be significantly underreported. In Japan,

283

Conservation and Management

for example, about 500 Asiatic black bears are killed annually by hunters compared with 1000–2000 kills of “nuisance” individuals (Oi & Yamazaki 2009). In Hokkaido, Japan, of 5668 brown bears recorded killed during 2010–2017 from an increasing population, 89% were non-hunting kills compared to 11% killed by hunters (unpublished data provided by T. Mano, Institute of Environmental Sciences, Sapporo, Japan). In contrast, in portions of Alaska with low human density, hunter kills predominate. “Defense of life or property” kills in Alaska were 5.2% of total reported kills of brown bears and 3.1% for American black bears (in urban areas 22.3% and 6.1%, respectively) (Miller & Tutterrow 1999). In Croatia, management plans anticipate that 80% of annual brown bear mortalities will come from hunting and 20% from accidents and nuisance bear kills (Huber et al. 2008a). Unreported bear kills occur everywhere and are thought to be high throughout the range of the world’s bears, especially in areas where bear parts, such as gall bladders and paws, are illegally traded (Servheen 1990; Gong & Harris 2006). Brown bears, American and Asiatic black bears, and polar bears are legally hunted in some areas. In these areas hunting is typically limited to sustainable levels of mortality and may focus on adult males. In terms of population growth rate, male mortalities have less impact than mortalities to adult females (e.g. Taylor et al. 1987; Harris et al. 2006). Non-sport kills, in contrast, are not as selective for larger individuals (males) (e.g. Huber et al. 2008b; Reljić et al. 2018). Hunting regulations protecting females accompanied by newborn cubs (or older offspring) also act to protect adult females. Sun bears may be hunted with special permits in Sarawak, Malaysia (Krishasamy & Shepherd 2014) but, like Andean, sloth, and panda bears, are not legally hunted elsewhere. Throughout the world where bears are hunted (excluding Alaska) the population management objectives are to allow hunting at levels that maintain or increase existing populations. Exceptions occur in some small areas where populations are sometimes managed to reduce conflicts with humans; information campaigns on methods to reduce conflicts and other management efforts can be effective at reducing unnecessary bear mortalities (e.g. Huber et al. 2008b; Proctor et al. 2018b). In the USA, brown bears currently can be hunted only in Alaska and American black bears can be hunted in all states with significant populations. In Canada, hunting of brown bears ceased in Alberta in 2006 to allow population recovery. Hunting in British Columbia was halted in 2017, although indigenous harvest is allowed. Canadian brown bear hunting occurs in Yukon Territory, Northwest Territories, and Nunavut, and American black bear hunting is allowed in all Provinces and Territories. In Eurasia, brown bears can be legally hunted in most areas with significant populations (e.g. Sweden, the Dinaric and Carpathian Mountains, Japan, and Russia) but not in areas with small populations including Spain, Italy, or where status is unknown such as China (Huber et al. 2008b; Chapron et al. 2014). In most of Russia where brown bears are hunted, such as in Karelia and Murmansk, about 10% of the

284

estimated population size is accounted to annual hunting rate, although in most years this quota is not achieved (Tirronen et al. 2015). In contrast to the rest of the world, in Alaska, both brown and black bears in most of the state are managed to reduce populations in the, as yet undocumented, expectation this will reduce predation on and increase hunter harvests of wild moose (Alces alces) and caribou (Rangifer tarandus) (Miller et al. 2017; Ripple et al. 2019). Fees paid by bear hunters may promote bear conservation through protection of habitat, hiring of bear management specialists to reduce conflicts and promote coexistence, and providing a constituency for sustainable management (e.g. Miller et al. 2013). In the Karelian part of Russia a license (state fee) costs about US$50 on state land, and on private land it could reach US$1000 or more (Danilov & Tirronen 2011). In many European countries, hunters pay the trophy value of a harvested brown bear (the value is higher for larger bears and can reach as high as 10,000 EUR). Within the European Union (EU), brown bear hunting is subject to rules established by the Habitat Directive. Most commonly, acceptable levels of harvest or mortality are established using estimates of population size combined with estimates of sustainable levels of mortality based on data on vital rates of productivity and mortality obtained from studies of radio-marked animals, hair snagging, mark– recapture, other available data, and modeling exercises (e.g. Miller 1990; Harris et al. 2006; Boulanger et al. 2008; McLellan et al. 2017; Reljić et al. 2018). Most harvest management schemes also involve efforts to document how many bears hunters in specific management areas legally kill. Sometimes collection of information on the sex and age composition of harvested individuals is also required, although this information must be used cautiously, as developed elsewhere in this chapter. Reported impacts of hunting on survivorship of cubs vary between North America and Europe. In Europe, studies conclude that male-biased hunting reduces brown bear cub and yearling survivorship as it results in more sexually selected infanticide by male bears who benefit by having the female come into estrus earlier than if her cubs survive (e.g. Swenson et al. 1997; Swenson 2003; Gosselin et al. 2014). In contrast, North America studies of brown bear and black bears conclude that hunting does not result in changes in cub survivorship or may increase survivorship if hunting and other mortalities result in a decline in male abundance or population density (Stringham 1980; Miller et al. 2003; Schwartz et al. 2003; McLellan 2005; Czetwertynski et al. 2007). Some jurisdictions also attempt to estimate mortalities that are not reported and incorporate these into mortality quotas. In both hunted and unhunted populations, unknown and unreported mortalities may represent a significant portion of total mortalities (Cherry et al. 2002; McLellan et al. 2018). In North America, only aboriginal groups have rights to hunt polar bears (Prestrud & Stirling 1994; Regehr et al. 2017). In the US, the Marine Mammal Protection Act (MMPA) of

Conservation and Management

1972 eliminated polar bear hunting by the general public. This Act authorized continued hunting by subsistence users resident in coastal areas that are dependent on such hunting for subsistence. Authorized sale of authentic articles of handicrafts and clothing made from polar bears by these subsistence hunters is allowed. There are no quotas unless a population is declared “depleted” under the MMPA. In Canada, Inuit and First Nations peoples are allowed to harvest polar bears and quotas are usually allocated to villages. Hunters may utilize the meat and, in some subpopulations, Inuit hunters may sell their rights to harvest polar bears to non-indigenous people who must partake of a sport hunt led by Inuit guides using traditional means. Hides can be sold commercially and most reach the international market. Quotas are established by research on polar bear abundance and trends. There is some controversy between Inuit hunters and scientists over population trends and the harvest quotas based on them. In both the US and Canada, the ultimate threat to polar bear populations is habitat loss caused by climate change, and not overharvest (see also Chapter 14).

Population and Trend Estimation for Bears There are no universal or widespread best ways to estimate bear abundance and trend across species and populations. Factors to be considered in selecting an approach include species and associated behaviors, habitat conditions, population density, and availability of financial and technical resources. Management needs for accuracy and precision vary widely between species and jurisdictions. In some cases available resources may be better spent in ways other than increasing the precision of population estimates, such as obtaining better estimates of the threats to populations and their habitats or controlling poaching. This is the situation for some populations of Asian and Andean bears where even the distribution of species may not be well known. In other cases, such as where American black bears are abundant, hunting quotas are set conservatively, habitat is not deteriorating, and management efforts to control unnecessary mortalities (such as to nuisance bears) are in place and successful, precise and large-scale population estimates may not be the best use of available funds. Accurate population estimates are commonly more critical for small and/or isolated populations, where concerns over loss of genetic diversity or population extinction exist. Although precision and accuracy in population estimation techniques have improved significantly in recent decades, all techniques have actual or potential biases and the presence and direction of these must inform decisions based on estimates obtained. In some cases, these biases can be minimized with the use of multiple sampling methods (Boulanger et al. 2008). Imprecision in estimates of abundance of hunted populations generally mandates setting harvest quotas conservatively. Assessments of abundance and trends based on impressions unaccompanied by rigorous measurements should be viewed cautiously (Garshelis & Hristienko 2006).

Many techniques used to estimate abundance in bear populations employ variations of capture–mark–recapture (CMR) techniques. CMR methods establish a known number of marked individuals randomly mixed into a population, followed by a “recapture” effort, which samples both marked and unmarked individuals. The population is estimated from the ratio of marked to unmarked individuals in the recapture sample(s). There are a number of options in how animals are “marked” and “recaptured” including de-facto “marks” based on DNA in hair or fecal samples (Woods et al. 1999; Bellemain et al. 2005), physical marks like tags or telemetry transmitters, chemical marking using tetracycline-laced baits, natural markings, and group characteristics documented by direct observations or remote cameras. Handling and fitting bears with radio transmitters and following them through time allows collection of data on age, births, deaths, movements, condition, habitat use, and other parameters that are usually extremely valuable to bear managers. Information on poor physical condition of handled animals can sometimes allow managers to anticipate declines in abundance prior to being able to detect this directly through abundance estimators (e.g. Stirling et al. 1999; Obbard et al. 2006; see also Chapter 14). In some cases, use of methods that do not involve handling bears may cost less, increase precision because of larger sample sizes, and better meet technique assumptions such as not creating trap-shy or trap-prone individuals. CMR approaches to estimating population abundance include a variety of assumptions and characteristics (Seber 1982; Cooch & White 2019). Closed models assume equal capture probability for all individuals, and demographic and geographic closure. Open population models do not assume closure, but require at least 3 years of data to estimate abundance based on year-specific estimates of apparent survival, capture probability, and number of new individuals (Cooch & White 2019). Spatial capture–recapture (SCR) models do not assume closure and directly estimate the area to which abundance estimates apply; this permits direct estimates of density (Royle et al. 2014; Hooker et al. 2015). In contrast to non-spatial approaches, SCR models can be used to examine variation in abundance/density within study areas (Kendall et al. 2019). Population densities are typically more valuable than just a population estimate, as they may reflect factors that vary spatially such as carrying capacity (food supply) or mortality rates. Also, estimates of density in small study areas may be extrapolated to other areas with comparable habitat and survival rates to obtain abundance estimates for larger, managerially significant geographic areas such as a park or reserve.

Techniques Requiring Handling of Bears Trapping and Tagging For many years the standard way to estimate abundance of American black bears was based on CMR techniques using

285

Conservation and Management

physically captured and marked bears followed by recapture effort(s) in a grid of baited traps/snares (examples cited in Pelton 2003). Costs associated with this technique typically limit its use to small geographic areas and, corresponding, small sample sizes. As with all CMR techniques, this approach has potential assumption violations that must be addressed including presence of trap-shy or trap-prone individuals and lack of geographic closure of the population (Seber 1982).

Radio Telemetry Most polar bear abundance estimates have involved multiyear CMR techniques (e.g. Derocher 2012). Radio-telemetry data and the Cormack–Jolly–Seber CMR models were used to estimate size and trend in various polar bear subpopulations (e.g. Amstrup et al. 2001; Stirling et al. 2011). In western Hudson Bay, Lunn et al. (2016) used live recaptures plus dead recoveries in a Bayesian implementation of multistate CMR models. Regehr et al. (2018) combined telemetry and CMR in an Integrated Population Model (IPM) for polar bears in the Chukchi Sea. In that study, multiyear capture data were used to estimate density and vital rates in a small intensively studied area with additional data added from radio telemetry. IPMs are powerful tools as they improve sampling efficiency/coverage and estimate precision by combining data from multiple sources. In Nepal, 17 radio-collared sloth bears were tracked for 2 years. Sloth bears are relatively social and sightings of nondependent unmarked bears with or near marked bears constituted the “recapture” sample (Garshelis et al. 1999a).

Known Fate Studies to Estimate Population Trend Population growth rate (λ) can be estimated with information on natality and survivorship obtained from observations of radio-marked bears (e.g. Eberhardt 2002). This approach has been used to estimate trend in a number of bear populations in North America (e.g. Garshelis et al. 2005; Schwartz et al. 2006) and Europe (Saether et al. 1998). When growth/decline rates based on known-fate telemetry studies are made for very large areas, it is important that telemetry collars be distributed geographically in proportion to bear abundance. This was the case for a study in a 24,000 km2 portion of the Northern Continental Divide Ecosystem (NCDE) (Mace et al. 2012). The annual minimum of 25 radio-collared bears were distributed within the NCDE based on detection results obtained from a DNA hair snaring study that applied equal sampling effort throughout the NCDE (Kendall et al. 2009, discussed below). This growth rate was applied to the initial DNA-based abundance estimate to derive annual estimates of abundance.

Using Offspring Survivorship Data to Estimate Trend Eberhardt (2002) proposed a sequence of changes in vital rates as population density for long-lived vertebrates increases, predicting an initial decline in survival of immatures, followed by an increase in the age of primiparity, a reduction in reproductive rates of adult females, and, lastly, increased mortality rates of adults. In the Yellowstone Ecosystem, a decline in cub and

286

yearling survival was observed in conjunction with a slowing of population growth (van Manen et al. 2016). This observation was interpreted as a possible expression of densitydependent regulation as this brown bear population neared its demographic maximal levels, or carrying capacity, in its core habitat (van Manen et al. 2016). Lower juvenile survival in unhunted National Park populations in Alaska compared to nearby hunted populations was also attributed to Park populations having reached their demographic carrying capacity while the hunted populations had not (Miller et al. 2003; Keay et al. 2018). In protected areas, adult males reach higher densities associated with higher survival, and intraspecific killing by adult males is a likely mechanism for lower survival rates of juveniles, particularly cubs. Thus, unhunted populations may approach their demographic carrying capacity at population numbers below the level that available food resources in their environment could sustain.

Unduplicated Counts of Females With Cubs to Estimate Abundance One of the longest continuous data sets for annual estimates of bear abundance is for brown bears in an isolated population occupying 23,833 km2 in the Greater Yellowstone Ecosystem, USA. Abundance in this area has been estimated since 1983 using unduplicated counts of females with cubs of the year (FFcoy) and a conservative rule set to avoid duplicate counting of groups of FFcoy (Schwartz et al. 2006; van Manen et al. 2018). Estimates of FFcoy are then extrapolated to obtain total population estimates. Similarly, FFcoy counts have been used since the beginning of the 1980s as an index for monitoring brown bear populations or estimating a minimum number of adult females in several small and medium-sized European populations (Ordiz et al. 2007).

Techniques Not Requiring Handling Bears DNA Sampling The development in the 1990s of laboratory methods that obtained useful genetic information from minute, degraded DNA samples from hair and feces (Taberlet & Bouvet 1992; Höss et al. 1992) revolutionized bear research and monitoring. Since the development of reliable techniques to identify bear species, sex, and individual identity from such samples (Taberlet et al. 1996), approaches involving sequencing of small DNA samples without the need to capture and handle individual bears have become widely used to estimate bear abundance (e.g. Woods et al. 1999; Bellemain et al. 2005; Proctor et al. 2010). Genetic detection generally samples a larger proportion of the population, resulting in more precise estimates of abundance/density than methods that involve handling. DNAbased techniques also provide information on effective population size (Pierson et al. 2018) and genetic diversity (Mikle et al. 2016). DNA from fecal samples is degraded, with lower genotyping success rates compared to hair samples (Paetkau 2003; Sorensen et al. 2017). Typically, 6–7 microsatellite loci and a sex primer are used to identify individuals. If the level of

Conservation and Management

genetic diversity in a population is unknown, it is necessary to conduct a pilot study prior to setting the number of loci that need to be examined to distinguish among individuals. Bear hair is typically obtained using hair corrals consisting of barbed wire encircling 3–6 trees positioned about 50 cm high with a scent lure at the center (Woods et al. 1999; Proctor et al. 2010; Figure 20.4). Other sampling techniques, such as hair from trees on which bears like to rub or trail hair-snares, do not require attractants and avoid potential behavioral effects on detection probability (e.g. Beier et al. 2005; Kendall et al. 2008). Scats can be obtained by systematic survey routes or grids, searches by scat-detecting dogs (Wasser et al. 2004), or volunteer collectors, such as hunters (Bellemain et al. 2005).

Sampling Hair to Estimate Abundance One of the largest-scale studies to estimate brown and black bear population abundance using DNA was conducted in the 31,410 km2 NCDE in western Montana including Glacier National Park (Kendall et al. 2009). Bears were sampled using two independent hair collection methods: hair corrals deployed on a grid and rub trees identified along trails, roads, and powerlines. Genotyping at seven loci identified 266 female and 182 male brown bears. The CMR model estimate of total population size was 765 (95% CI = 715–831) or 2.4/100 km2. Obtaining hair samples from multiple methods (e.g. corrals and rubs) produces more precise and less-biased estimates than from either method alone (Boulanger et al. 2008). The same NCDE brown bear hair corrals and rub trees simultaneously collected black bear hair. This was used to estimate abundance and density in a sympatric American black bear population in Glacier National Park (4100 km2) using closed population CMR models (Stetz et al. 2014). Based on genotyping at six loci, density was estimated at 11.4 bears/100 km2 (95% CI = 9.9–13.0). DNA hair collection using hair corrals in small study areas resulted in density estimates for Andean bears and Asiatic black bears. For Andean bears, DNA analyses have been attempted (e.g. Viteri 2007) but require more development (Garshelis 2011). DNA techniques have also been attempted for Asiatic black bears in India (Mukesh et al. 2015). The brown bear population in the Gobi Desert, Mongolia was sampled using hair corrals surrounding 13 feeding stations at most water sources used by the population (Tumendemberel et al. 2015). Of the 600 hair samples collected in 2009, 205 were genotyped to 21 individuals. Genetic variability was low compared to other brown bear populations, so samples were genotyped at 12 microsatellite loci to ensure reliable identification of individuals. Using CMR models, population abundance exposed to the hair corrals was estimated to be 22 bears (95% CI = 21–29).

Clustered Hair Sampling Humm et al. (2017) estimated density of the five major American black bear subpopulations in Florida, USA, with SCR models using genetic detection data from clustered hair corrals. The efficiency of the clustered sampling design coupled

with the use of several landscape covariates allowed researchers to estimate abundance (3916 bears, 95% CI = 2914–5451) in a 38,960 km2 study area.

Sampling Feces to Estimate Abundance Abundance estimates using DNA from feces have been obtained for brown bear populations in Eurasia. For example, in Sweden the brown bear population was estimated using CMR methods at 3298 (95% CI = 2968–3667) using six microsatellite loci that identified 1358 individuals (Bellemain et al. 2005; Kindberg et al. 2011). DNA in fecal samples was used to estimate brown bear numbers and evaluate effective population size in Slovenia and Croatia (Skrbinšek et al. 2018). Four abundance estimates in China (1974–1977, 1985–1988, 2000–2004, 2011–2014) were based on measurements of bamboo segments (assumed to represent individually specific bite sizes) found in feces. These results concluded there was a gradual national increase in giant panda abundance (Wei et al. 2018). Another approach, based on DNA found in panda feces in one area, resulted in a nationwide estimate that approximately doubled the bite size estimate (assuming the density relationship held across the panda range) (Zhan et al. 2006). The DNA-based results were considered to be biased high, because of closure bias and other considerations (Garshelis et al. 2008; see also Zhan et al. 2009).

Aerial Surveys Aerial survey methods are most effective in open habitats where bears are readily visible. Polar bear abundance in Hudson Bay, Canada, was estimated using aerial double-observer techniques based on CMR principles (Stapleton et al. 2014, 2015). This involved independent observers in the front and rear of an aircraft; a polar bear was considered “marked” if seen by the forward observer and “recaptured” if also seen by the other observer. Transects were flown during the summer ice-free season when all polar bears are on land. A similar approach was attempted for brown bears in Alaska but had low precision (Walsh et al. 2010). Aerial search techniques similar to those of Miller et al. (1997) but that did not require radio-collared animals were used to estimate brown bear abundance on 19,988 km2 of treeless habitat in north-western Alaska (Schmidt et al. 2017a). Initial “marking” was based on the spatial location and characteristics (e.g. coloration, group size) of each group obtained and photographed during aerial searches. The “recapture” sample was acquired during independent aerial searches where location and these characteristics were used to determine if bears seen were “recaptures” or previously unseen bears. The cost of this method was ~10% of the cost of smaller area density estimates using radio-collared bears (Miller et al. 1997). Densities from each technique were broadly comparable, but the approach without telemetry had lower precision. For polar bears and brown bears, double-observer aerial surveys are sometimes conducted simultaneously with distance sampling techniques (e.g. Crête et al. 1991; Stapleton et al.

287

Conservation and Management

2015; Obbard et al. 2015; Schmidt et al. 2017b). Distance sampling techniques are based on assumptions involving models of relative sightability of bears occurring at different distances from transect lines flown by aircraft. For polar bears, transects were perpendicular to the shore and transect spacing was based on stratification of the survey area based on preexisting knowledge of bear density (Stapleton et al. 2015). An evaluation of approaches involving distance sampling alone and combined with CMR double-observer data concluded that distance sampling alone provided better results with less complexity (Schmidt et al. 2017b).

Combining DNA with Telemetry Data to Estimate Abundance at Edge of Range Jurisdictions interested in estimating population size at the edge of its range need to account for part-time occupancy by transboundary bears. Norway has politically established goals for reproduction of brown bears in five management areas that occur at the edges of the bears’ ranges in adjacent countries. The Norwegian goal is 13 annual reproductions in these five areas. Annual counts of individuals based on DNA gathered from scats and hair and telemetry-derived knowledge about demography, home range size and use provide annual management area-specific estimates of reproduction that can be credited to Norway (Bishof & Swenson 2012).

Using DNA to Estimate Trend Genetic detection data were used to estimate brown bear population trend and local density based on hair samples collected at 3579–4802 natural bear rub sites along trails, roads, and fence and utility lines in western Montana’s NCDE (Kendall et al. 2019). Rubs were visited 2–4 times annually during 2004 and again during 2009–2012, resulting in detection of 249–355 individual bears each year. SCR models were used to estimate an annual growth rate of 1.043 (95% CI = 1.017–1.069). During the study, the highest population densities shifted from primarily in the north to a more even distribution throughout the NCDE. Simulation studies indicated that rub tree surveys were likely to be a reliable way to detect population trends (Stetz et al. 2010). DNA-based approaches combined with SCR models promise to transform trend monitoring for bears and other difficult-to-study species across large landscapes. Two abundance estimates conducted 10 years apart were used to estimate brown bear population trend in the Canadian Rocky Mountains (Stenhouse et al. 2015). The estimates were made with samples from hair corrals, genotyped at seven microsatellite loci, and detection data analyzed using SCR models. The population was estimated to have grown at an annual rate of λ = 1.07, but there was a large amount of uncertainty (95% CI = 0.84–1.30). Nonetheless, this study concluded that spatially explicit methods obtained robust abundance estimates with less sampling effort than traditional CMR models by enabling the use of sampling stratification, thereby reducing project costs.

288

Marking Bears With Ingested Tetracycline to Estimate Abundance The number of black bears in Minnesota and parts of Michigan (USA) was estimated using a CMR estimator by hanging widely spaced baits laced with tetracycline in trees and determining how many were ingested by bears (Garshelis & Visser 1997). Tetracycline was identified by laboratory analysis in the teeth and bones of hunter-killed bears that ate these baits. This study resulted in estimates of approximately 15,600 black bears (± 1500) in Minnesota and 6600 (± 1400) in Michigan. A similar tetracycline study was also done in Michigan’s Upper Peninsula (Belant et al. 2011).

Estimating Abundance and Trend Using Remote Cameras Motion-sensitive cameras are used to detect and identify images of many wildlife species (Burton et al. 2015). Camerabased CMR studies using distinctive chest patterns to identify individuals were used to estimate density of Asiatic black bears and sympatric sun bears in two study areas in Kha Yai National Park, Thailand (Ngoprasert et al. 2012, 2015). Imprecise density estimates for Asiatic black bears were 4.3–5.9/100 km2 and 8–29/100 km2 for sympatric sun bears. Camera traps were used to estimate Andean bear density in a Bolivian pilot study (Rios-Uzeda & Wallace 2007). In a potentially useful advance, machine learning techniques (“deep learning”) identified faces of individual brown bears in three study areas based on camera trap pictures with a claimed 93% (SE = ± 4.9%) accuracy (Clapham et al. 2018).

Lower Technology or Cost Methods Techniques used to determine trend in bear species/populations where resources are most limited include expert opinion, surveys that enumerate bear sign, changes of habitat extent, and perceived levels of hunting pressure. It is not possible to estimate confidence intervals around trend estimates made using these methods. For example, results from several national Andean bear assessments based on an assortment of techniques, including genetics, ecological modeling, and sign surveys, were combined to give a crude range-wide estimate of 13,000–18,000 bears (5–7/100 km2) over the species range of 260,000 km2 (Velez-Liendo & Garcia-Rangel 2017).

Occupancy Changes in presence of bears can be used in occupancy models to estimate increases or decreases in abundance in space or time. Presence data can come from observation (live capture, direct sightings of bears or tracks, remote camera images, or local knowledge/expert opinion) or DNA samples from hair, scats, or other sources. Occupancy by brown bears in formerly vacant habitats adjacent to growing populations in the NCDE, Greater Yellowstone Ecosystem, and Sweden was consistent with other studies, indicating growing populations expanding into former range (e.g. Schwartz et al. 2002; Kendall et al. 2019). Population trends for sun bears were measured in Sumatra using camera trap detections in an occupancy-based

Conservation and Management

sampling framework (Wong et al. 2013). A trend estimate for sun bears across their global range in South-East Asia was based on the relationship between tree cover and relative density of sun bears to project population change due to habitat loss (Scotson et al. 2017a). In Borneo, sun bear presence documented with bear DNA obtained from leeches (Haemadipsa spp.) correlated with presence data from camera traps (Abrams et al. 2019). Because occupancy determination only requires species identification, DNA laboratory analysis is inexpensive compared to methods requiring identification of individuals. In some portions of Asia where distribution of bears remains uncertain, useful information on occupancy as well as subjective information on trend, threats and other issues have been gathered by a combination of questionnaires and sign surveys. Typically, this involves creation of a grid of cells covering a province (or other pertinent area), asking residents of grid cells about bear occupancy, trends, and problems and then asking residents to show investigators areas of bear activity where sign can be seen. This approach has been used for Asiatic black bears in south-western China (Liu et al. 2011) and for sloth bears in Sri Lanka (Ratnayake et al. 2007). In places where Asiatic black bears and sun bears are sympatric, claw marks on trees were used to differentiate between these species (Steinmetz & Garshelis 2008). Population estimates for Asiatic black bears and sun bears were attempted using questionnaires that asked regional experts for their subjective estimates of park populations with perceived confidence ranges (Crudge et al. 2019). For both species, the IUCN Bear Specialist Group estimated a rangewide percent decline over 30-year time windows, weighted by extent of occurrence within range. This method suggested a 30% decline of Asiatic black bears and up to 40% decline of sun bears (Garshelis & Steinmetz 2017; Scotson et al. 2017b). In India, a country-wide sloth bear population estimate of 6000–11,000 was based on questionnaire data sent to 800 wildlife researchers, managers, and naturalists, in combination with satellite-based habitat cover maps (Yoganand et al. 2006). Questionnaire surveys indicated a reduction of brown bear distribution in Pakistan of 15.7% since the 1950s (Abbas et al. 2015). In Indonesian Borneo (Kalimantan), sign transects were used to monitor trend in sun bears during 2000–2010 in an area of regenerating burned forest. For Andean bears in Colombia, efforts are ongoing to monitor trend from presence data based on bear sign observed along non-random line transects (Acevedo et al. 2017).

Hunter Observations or Kill Data Estimates of population trend can be obtained in jurisdictions where hunters are well organized and will report observations of bears and hours afield. In Sweden, about 200,000 hunters report the number of hours hunted and number of bears observed during the first seven days of the moose hunting season. Kindberg et al. (2009) found that the resulting bear

density index documented bear distribution and showed a strong linear correlation with DNA-based population estimates at the wildlife management unit scale (ca. 1000–2500 km2). Kindberg et al. (2011) combined these trend estimates with periodic DNA-based regional population estimates to obtain an estimate of brown bears in Sweden. In several European countries where bears are hunted, such as Slovenia, Croatia, and Romania, there are permanent feeding sites equipped with observation/shooting towers where hunters make visual counts of bears seen on predetermined nights. The standardized reports are summed and, for decades, were used to estimate the population size and decide the hunting quotas (Huber et al. 2008a). More recently, the same data were used only to evaluate population trend. This supplements more accurate CMR estimates based on genetic detection, which is too expensive to be done each year. A weak point of counts at feeding sites is that annual fluctuations in hard or soft mast production can change bear movement patterns (Ryan et al. 2007) and encounter frequency with feeding sites. Thus counts at feeding sites can vary in response to changes in mast production rather than changes in population numbers. Jurisdictions where hunters are required to submit teeth from harvested bears for age determination and allow managers to inspect hides to determine gender expect that these data can be used to assess population trends (e.g. Kolenosky 1986). For bears, there is not yet a clear-cut way to interpret these data, although Population Reconstruction Analysis (Millspaugh et al. 2010; Skalski et al. 2012) may be useful in some circumstances. Harvest data are difficult to use to infer trend because of frequent and difficult-to-quantify biases in hunterharvests associated with variable or changing vulnerabilities of sex/age categories (e.g. Garshelis 2002; Hatter et al. 2018; Newton & Obbard 2018), sometimes associated with changes in hunting regulations. Regardless, when associated with other information, these data can sometimes prove helpful in estimating trend in bear populations over large geographic scales (Fieberg et al. 2010; Allen et al. 2018) or can flag conditions where more intensive studies on abundance or trend may be needed.

The Management of Human–Bear Conflict Human–bear conflict (HBC) is a topic that is becoming increasingly important for the conservation and management of bears throughout the world (Ambarlı & Bilgin 2008; Can et al. 2014; see also Chapter 15). A good definition of HBC is provided by the Human–Bear Conflicts Expert Team (part of the IUCN Bear Specialist Group): “HBC is any situation where wild bears undesirably use or damage human property, where wild bears harm people, or where people perceive bears to be a direct threat to their property or safety.” An important concept in the discussion of human–bear conflict is the notion of an affected stakeholder, which we define as those people living in a landscape with bears and that are impacted by HBC. Affected stakeholders are

289

Conservation and Management

distinguished from non-affected stakeholders (e.g. people in conservation organizations) that may be invested in having bears on the landscape but are not impacted directly from their activity. Identifying and understanding impacts to affected stakeholders is important because failing to address their concerns can lead to a variety of sociopolitical impacts influencing the potential for bears and humans to coexist. Three general strategies are used (often in conjunction) to minimize HBC: managing anthropogenic resources, managing bears, and compensating for HBC impact. Managing anthropogenic resources usually, but not always, involves efforts to reduce the availability or access to human resources. For example, reduction of anthropogenic resources formed the foundation of reducing HBC for brown bears and black bears in National Parks throughout the US and Canada by reducing the availability of garbage and other anthropogenic food in the Parks (e.g. Denali National Park 2003; Glacier National Park 2010). Applying the principle of anthropogenic food reduction to other systems (e.g. livestock spread across a landscape or agricultural crops in mixed landscapes) can be more difficult because authority to strictly manage these resources, particularly on private lands, can be limited and costs for doing so can be prohibitive. Instead of reducing anthropogenic resources, in some situations the opposite approach is taken by providing additional food sources (i.e. supplemental feeding) as a means of reducing HBC. For example, in the Black Sea forests of Turkey the decline of wild orchard trees resulted in increased conflict by brown bears. In response, the Turkish Department of Forestry planted orchard trees while implementing regeneration plantations to help reduce HBC. In the north-western US, timber industries created feeding stations to feed black bears in an effort to reduce bears feeding on the inner cambium of newly planted conifer trees (Ziegltrum 2004). The science supporting this type of effort is mixed and generally most management authorities are not in favor of supplemental feeding. Perhaps the most effective tool for preventing conflict is electric fences, and this tool has been used effectively in a wide variety of contexts to protect agricultural crops and livestock but can be expensive to install and operate. In general, nonlethal tools become less effective as anthropogenic resources comprise a greater percentage of the landscape or are more dispersed. When non-lethal methods are ineffective it sometimes becomes necessary to lethally remove bears. In developed countries, such control actions are generally performed by management authorities, whereas in developing countries management authorities are often not available to carry out such operations and thus locals may take it upon themselves to carry out lethal control actions. In some situations, sport hunting is promoted as a method for reducing HBC. However, there is little evidence that sport hunting reduces HBC (Obbard et al. 2014) unless hunting is so excessive that it significantly reduces the bear population to very low densities.

290

Another strategy for dealing with HBC is paying compensation for the economic damage to affected stakeholders. For example, Turkey established an agricultural insurance pool where all damage caused by wildlife can be compensated, if the farmer pays half of the small cost of insurance (US$2 for a sheep, US$5 dollar for a cow or beehive per year) (Tarsim 2018). In Sweden, concerted effort has gone into recovering their brown bear population (Swenson et al. 1995). In areas where recovery has occurred, an important affected stakeholder group are the indigenous Sami herders that utilize landscapes in Scandinavia to maintain the traditional culture of herding semidomestic reindeer. Brown bears prey on reindeer, resulting in negative economic impacts to the Sami herders. To minimize the impacts, national policy requires that herders are compensated for losses to bears and it is believed that this compensation is an important component of a larger management scheme that helps maintain tolerance for bears. In the lower 48 United States, grizzly bears have been recovered in critical components of their historic range (Gunther 2015). A major part of the recovery program for the grizzly bear was assuring a clear and active response to HBC (primarily predation on livestock, impacts to beehives, and breaking into houses) (Interagency Grizzly Bear Guidelines: Mealy 1979). Agencies, in conjunction with the enhanced efforts of private groups, worked together to develop and implement a comprehensive and responsive management system, assist affected stakeholders in preventing conflicts, and increased the non-lethal management of conflict bears. Through these kinds of efforts (Proctor et al. 2018b), HBC was minimized and the number of bears that had to be lethally removed was considerably reduced, thus promoting the growth and recovery of the bear population. Many of the management examples we present have occurred in developed countries with abundant resources available to manage conflict. These types of resources are generally not available in developing countries, which begs the question of whether the management of HBC is a conservation luxury of developed countries. For example, throughout South-East Asia, conversion of native forest for agriculture has helped cause precipitous declines for a number of bear species including the sloth bear, Asiatic black bear, and sun bear (Scotson et al. 2014; Wong et al. 2015; Debata et al. 2016). The increased presence of people and impact on the landscape has created dynamics where bears are coming into greater contact with people, which has resulted in higher rates of attacks on people and greater utilization of human food crops. Retaliatory killing of bears in an effort to reduce HBC is common and contributes to the reduction in the bear population. Compounding this problem are the economic incentives associated with poaching for sale of bear parts. In many cases, killing bears to protect crops may be a cover for the true desire to kill bears to sell their body parts. Critical to the successful management of HBC is detailed understanding of human culture where bears and humans interact and integrating

Conservation and Management

affected stakeholders into the process of developing solutions to manage HBC (Redpath et al. 2017).

Managing the Trade in Bear Parts Of the world’s eight extant bear species, three species, Asiatic black bears (Ursus thibetanus), sun bears, and brown bears, are most threatened by the illegal wildlife trade. All are threatened by poaching and are the most commonly seized species of bears across Asia (Burgess et al. 2014). Bears are poached mainly for their gall bladders, but also for bear cubs, paws, claws, and other body parts. Bile from bear gall has been used in Chinese traditional medicine for thousands of years as a treatment for a wide range of inflammatory and degenerative ailments (Feng et al. 2009; Li et al. 2016). Traditionally, bear bile was sourced from the gall bladders of bears hunted and killed in the wild. Since the 1980s, several Asian countries have begun keeping wild-caught bears in captivity and also farming live bears, predominantly Asiatic black bears, for their bile (Servheen & Mills 1991; Servheen et al. 1999). China is the leading consumer and producer of bear bile tapped from captive bears, with around 20,000 bears on socalled farms. Captive (often wild-caught) bears are also kept for commercial bile extraction in South Korea and Vietnam, with fewer in Lao PDR and Myanmar (Animals Asia 2011; BANCA 2017; Willcox et al. 2016; World Animal Protection 2017; Livingstone et al. 2018). Zoos and private collections also play a role in illegal bear trade. For example, Japanese “bear parks” hold more than 1000 bears of various species and some of these “parks” trade in bear bile products obtained from their captive bears (Mills & Servheen 1991; Togawa & Sakamoto 2002; Foley et al. 2011). In Lao PDR, a number of private collectors, including Chinese-owned “zoos,” hold Asiatic black bears and are suspected to be involved in the illegal trade of live bears (EIA 2015; Livingstone et al. 2018). International trade in bears and bear parts is regulated by the Convention on International Trade in Endangered Species (CITES). Asiatic black bears and sun bears are classed as CITES Appendix I, which restricts international trade in these species, unless under exceptional circumstances. Import permits can be issued if specimens are not to be used for commercial purposes, or if the import purposes are not deemed to be detrimental to the survival of the species. There may be exceptions made for specimens bred in captivity (CITES 2000). Brown bears are Appendix II, in much of their range, except for populations in Bhutan, China, Mexico (extirpated), and Mongolia, which are covered by Appendix I (www.cites.org). Differences in Appendix classifications between and within bear species create difficulties in controlling illegal trade, due to difficulties in identifying species by parts of the animals and the origin of products being traded (CITES 2000). Domestic trade in farmed bear products is permitted within China, Myanmar, and Lao PDR, providing that products come from second-generation captive-bred bears. This provision allows the promotion of bile sales for commercial purposes and

creates a trade loophole due to deficiencies in monitoring the origin of captive bears and problems differentiating parts from wild versus captive-bred bears. The impact that illegal trade in bears has on wild bear populations has not been quantified, and assessments are largely speculative (Garshelis 1997; Garshelis & Steinmetz 2017; McLellan et al. 2016; Scotson et al. 2017a), but many forest areas in South-East Asia are becoming devoid of bears and other widllife, strongly indicating that there is a poaching crisis (Harrison 2011; Gray et al. 2017). It is widely assumed by conservation scientists that bear farming has a negative impact on wild bear populations because (1) killing of wild bears for trade of their parts is unregulated and largely unknown; and (2) there is a two-tiered market for both farmed product and a more desirable wild product (Dutton et al., 2011; IUCN 2012; Livingstone & Shepherd 2014; Garshelis & Steinmetz 2017; Scotson et al. 2017a). There is need for a clear and united strategy among organizations invested in conservation of bears to reduce consumer demand for both farmed and wild bear products. Monitoring of the trade in bear parts would benefit from a central database devoted to the bear trade, alongside enhanced communication between organizations working with bears to optimize opportunities to collate data in a meaningful way (Burgess et al. 2014). Understanding consumer motivation, promoting alternatives to bear bile in traditional medicine, and educating and changing the behavior of potential consumers of bear products could be effective tools to reduce bile trade. There have been some successes in reducing the bile trade evident in Vietnam and South Korea, where bear farming is being phased out (Willcox et al. 2016; World Animal Protection 2017). In summary, the trade in bears and bear parts continues to be a serious problem for the conservation of Asian bears. Despite more than 25 years of effort to reduce the demand for bear bile and to decrease the number of wild bears killed for the sale of their parts, the trade continues and may be increasing in some countries.

Artificial Feeding and Bear Conservation: Advantages and Disadvantages

“Artificial feeding” of bears includes all efforts to deliberately provide food (other than garbage) in natural habitats as a bear management tool. Objectives can include baiting for hunting, diversionary feeding to decrease human–bear conflicts, concentration of bears for viewing or photography for human recreation, and/or nutritional enhancement of populations. Artificial feeding has been a tool used to promote bear conservation, particularly for threatened populations, such as the feeding program for the Gobi brown bear in Mongolia. Bear feeding can sometimes have unintended negative consequences for species and ecosystems (reviewed by Penteriani et al. 2010, 2017), such as alterations of social and trophic interactions; changed behavior, activity, and movement

291

Conservation and Management

patterns (Krofel et al. 2017); changes in reproduction; and pathogen transmission. Therefore, bear feeding is of conservation concern and should be evaluated on a case-by-case basis if it is to continue (Skuban et al. 2016; Krofel et al. 2017; Penteriani et al. 2017). Artificial feeding of bears is a common practice in many European countries. Increasing bear availability to hunters is the main motivation for such feeding in countries where bears are hunted, followed by increased reproduction and body size for hunted bears. Artificial feeding of bears occurs year-round in many cases and usually occurs at established feeding sites. Artificial foods for bears are also used by many non-target species of birds and mammals. Currently, more than 80% of the European countries where bears occur feed them intentionally or unintentionally. Bear feeding takes place in all European countries where brown bears occur except Albania, Andorra, Belarus, Greece, Italy, Spain, and Liechtenstein. Corn and livestock remains (whole carcasses and slaughter remains) are the most common foods used to feed bears in Europe. More than 60% of the areas where targeted bear feeding occurs have regulations regarding the type and amount of food that can be provided and the periods when feeding can occur, although often such regulations are not fully implemented (Huber et al. 2008b). In most cases, bear feeding has not been subject to an appropriate impact assessment, even if it occurs in Natura 2000 sites (European Union 1992), Emerald Network areas, or other protected areas. In both Emerald and Natura 2000 sites, national authorities have a legal obligation to ensure that feeding is in conformity with the ecological requirements of bears and other species for which the sites are designated, and that any significant adverse effects on these species are avoided. Unintentional feeding of bears where large herbivores are fed should be minimized where possible (Selva et al. 2014). The Large Carnivore Initiative for Europe (LCIE) completed a study on artificial feeding of bears as a management tool, which was submitted to the Council of Europe where it was adopted as a Recommendation by the Bern Convention (Convention on the Conservation of European Wildlife and Natural Habitats 2018). This report noted that there is an urgent need to reevaluate the practice of artificial feeding of bears from cultural, ecological, conservation, and legal perspectives. While recognizing that in some situations it can be an appropriate management tool in general, the LCIE recommended against artificial feeding of bears and would like to see a progressive decrease in the practice. In areas where it is currently practiced, the LCIE recommends a detailed case-bycase evaluation of the intended goals and the potential impacts on target and non-target species, as well as wider ecosystem effects. There is also a need to develop and enforce clear regulations to govern the practice that relate to periods of the year when feeding should be conducted, the location of feeding sites, and the type and amounts of food that can be used.

292

Critical Needs for the Conservation of Bears and Bear Habitat in South-East Asia South-East Asia’s bears would benefit most from two key management actions: (1) reducing habitat loss, particularly deforestation by timber harvest and conversion of forests to plantation agriculture; and (2) reducing illegal poaching and killing bears for the sale of their body parts. Neither action is easily achieved under the current climate of increasing and expanding human populations, minimal political support for bear conservation, and few financial resources for bear research, monitoring, and management. Deforestation rates are highest in Myanmar, Indonesia, and Malaysia (Scotson et al. 2017b). Bear populations are negatively impacted by extensive clear-cutting for agricultural plantations (e.g. oil palm, rubber), unsustainable logging (both legal and illegal), and forest fires (Brown et al. 2005; Meijaard et al. 2005; Tumbelaka & Fredriksson 2006; Wong & Linkie 2013). Indonesia and Malaysia are both leaders in global oil palm production, which is a primary driver of bear habitat destruction (Miettinen et al. 2011; Margono et al. 2014). In Indonesia, 40% of the forest lost between 2000 and 2012 was lost in areas where logging was restricted or prohibited, including within protected forests due to deficiencies in governance (Murdiyarso et al. 2011; Margono et al. 2014). In Kalimantan alone, 56% of protected lowland forests were cleared from 1985 to 2001 (Curran et al. 2004). A priority for bear conservation is to protect high conservation value forests from conversion to other land uses, to eliminate unsustainable logging, and prevent forest fires. Similarly, managers should prioritize restoration of degraded forest in human-modified landscapes through reforestation, connectivity management, and enhanced protected status, and enforcing buffer zones to prevent further agricultural expansion into surrounding protected areas. Ultimately, reducing the trade in bear parts would be one of the most highly beneficial steps for the persistence and recovery of Asian bears throughout their range, especially as trade is increasingly moving toward the last strongholds in Asia (Shepherd & Shepherd 2010; Krishnasamy & Shepherd 2014). Asian bears are legally protected from hunting and trade throughout most of their range, but in many cases this is paper protection only. Deficiencies in law enforcement are recognized as the major ongoing weakness in bear conservation throughout Asia (Burgess et al. 2014; Baskin 2016). There have been a few successes, for instance in Cambodia where dedicated agencies operate with steady technical and funding support (e.g. Broadis 2011). Reduction of bear mortality by clearing forests of snares is urgently needed throughout much of the remaining bear range throughout Asia, and long-term measures are needed to prevent the problem recurring in cleared habitats (Gray et al. 2017).

Conservation and Management

Intensive Management of Brown Bears and Black Bears in Alaska Attitudes toward hunting brown bears in North America vary from almost no hunting of a mostly healthy population in British Columbia (Canada) to hunting of a population recently delisted from Endangered Species Act protections (Wyoming and Idaho, USA – hunting put on hold in 2018 by a court decision), to widespread efforts in Alaska (USA) designed to intentionally reduce populations of both black and brown bears. The Alaskan efforts are mandated by a state “Intensive Management” law (AS 16.05.255, 1994) requiring wildlife managers to reduce the abundance of bears and wolves (Canis lupus) in hopes this will result in increased hunter harvests of wild ungulates such as moose, caribou, and deer (Odocoileus hemonius sitkensis) (Miller et al. 2017; Ripple et al. 2019). Efforts to reduce bears involve liberalized hunting regulations including all-year hunting seasons, individual annual hunting quotas of two per year for brown bears and up to five per year for black bears, allowing hunters to sell the hides and skulls of bears they shoot, elimination of requirements to purchase a tag prior to hunting, allowing hunters to bait both species of bears, allowing take of bears using snares, and relaxation of rules regarding use of aircraft for hunting. This has resulted in dramatic increases in numbers of bears harvested without corresponding efforts to document trends in bear populations (Miller et al. 2017; Ripple et al. 2019).

The Future for Bears Around the World In this chapter we have reviewed the major conservation issues bears face and highlighted management actions that can

References Abbas, F., Bhatti, Z., Haider, J. & Mian, A. (2015). Bears in Pakistan: distribution, population biology and human conflicts. Journal of Bioresource Management 2(2): 1–13. Abrams, J. F., Hörig, L., Brozovic, R., Axtner, J., et al. (2019). Shifting up a gear with iDNA: from mammal detection events to standardized surveys. Journal of Applied Ecology 56: 1637–1648. Acevedo, C., Bernal, N., Bianchi, G., et al. (2017). Conservamos la vida: Andean bear conservation at the landscape scale. In: Molina, S., Zug, B., Velez-Liendo, X., et al. (Eds.), 25th Conference on Bear Research and Management (p. 9). Quito, Ecuador: International Association for Bear Research and Management. Allen, M. L., Norton, A. S., Stauffer, G., et al. (2018). A Bayesian state–space model using age-at-harvest data for estimating the population of black bears (Ursus

address these conservation issues. The future of bears across the world is bright for some species but dark for others. In some areas such as North America (North American black bear and brown bear) and in parts of Europe and Asia (brown bear), bear populations have increased and stabilized because of increased management effort and increasing support for bears and their needs by the humans who share habitat with them. However, for most bear species, the future is uncertain and increasingly dire. Andean bears continue to be threatened by habitat loss and human encroachment. In much of Asia outside Japan, Asiatic black bear, sloth bear, and sun bear populations are increasingly threatened by unmanaged excessive mortality combined with habitat loss to timber harvest, plantation agriculture, and human encroachment. The longterm future for polar bears is threatened by the unmanageable threat of climate change that will irrevocably change the unique habitat on which they depend. Giant pandas are fragmented into small populations despite intense conservation efforts. There have been advances in conservation techniques for bears, but the application of these techniques is often difficult due to lack of funding and opposition to bears by humans who are sometimes in conflict with them. The primary determinant whether bears expand and flourish or decline and disappear is not only science or the commitment of those who seek to help bears, but also the attitudes and support (or lack of support) of the people who live, work, and recreate in bear habitat and the political forces that represent these local people. Improving public and political support for bears is the most important need if we are to realize successful bear conservation and management.

americanus) in Wisconsin. Scientific Reports 8: 1–12. Ambarlı, H. & Bilgin, C. C. (2008). Human–brown bear conflicts in Artvin, northeastern Turkey: encounters, damage, and attitudes. Ursus 19(2): 146–153. Amstrup, S. C., McDonald, T. L. & Stirling, I.I. (2001). Polar bears in the Beaufort Sea: a 30-year mark–recapture case history. Journal of Agricultural, Biological, and Environmental Statistics 6: 221–234. Animals Asia. (2011). Bear farming industry in China. Chengdu, China: Animals Asia Foundation. Atwood, T. C., Peacock, E., McKinney, M. A., et al. (2016). Rapid environmental change drives increased land use by an arctic marine predator. PLoS ONE 11(6): e0155932. https://doi.org/10.1371/journal .pone.0155932 BANCA. (2017). The bear trade and bear farms in transboundary regions regions

in Eastern Myanmar. Yangon, Myanmar: Biodiversity and Nature Conservation Association. Available from www.banca-env.org/wp-content/ uploads/2018/05/the-bear-trade-andbear-farms.pdf Bargali, H. S., Akhtar, N. & Chauhan, M. P. S. (2005). Characteristics of sloth bear attacks and human casualties in North Balispur Forest Division, Chhattisgarh, India. Ursus 16: 263–267. Bargali, H.S., Akhtar, N. & Chauhan, M. P. S. (2012). The sloth bear activity and movement in highly fragmented and disturbed habitat in central India. World Journal of India 7: 312–319. Baskin, K. (2016). Hunting as sustainable wildlife management. Mammal Study 41: 173–180. https://doi.org/10.3106/041 .0402. Beier, L. R., Lewis, S. B., Flynn, R. W., et al. (2005). A single-catch snare to collect brown bear hair for genetic

293

Conservation and Management

mark–recapture studies. Wildlife Society Bulletin 33(2): 766–773. Belant, J. L., Etter, D. R., Mayhew, S. L., Visser, L. G. & Friedrich, P. D. (2011). Improving large-scale mark–recapture estimates for American black bear populations. Ursus 22(1): 9–23. Bellemain, E., Swenson, J. E., Tallon, D., et al. (2005). Estimating population size of elusive animals with DNA from huntercollected feces: four methods for brown bears. Conservation Biology 19(1): 150–161. Berkes, F. (2007). Community-based conservation in a globalized world. Proceedings of the National Academy of Sciences 104(39): 15,188–15,193. http://dx .doi.org/10.1073/pnas.0702098104. Bischof, R. & Swenson, J. E. (2012). Linking noninvasive genetic sampling and traditional monitoring to aid management of a trans-border carnivore population. Ecological Applications 22: 361–373. Boulanger, J. & Stenhouse, G. B. (2014). The impact of roads on the demography of grizzly bears in Alberta. PLoS ONE 9(12): e115535. https://doi.org/10.1371/journal .pone.0115535 Boulanger, J., Kendall, K. C., Stetz, J. B., et al. (2008). Use of multiple data sources to improve DNA-based mark–recapture population estimates of grizzly bears. Ecological Applications 18(3): 577–589. Broadis, N. (2011). Evolving threats; lessons learnt from 15 years of bear protection in Cambodia. Oral presentation, 21st International Conference for Bear Research and Management, Delhi, India. Brown, T. H., Simangunsong, B. C. H., Sukadri, D., et al. (2005). Restructuring and revitalization of Indonesia’s woodbased industry: Synthesis of three major studies. Jakarta: MoF-NRM-CIFORMFP-Forest Trends-ITTO. Burgess, E. A., Stoner, S. S. & Foley, K. E. (2014). Brought to bear: An analysis of seizures across Asia (2000–2011). Petaling Jaya, Selangor, Malaysia: TRAFFIC Southeast Asia. Burton, A. C., Neilson, E., Moreira, D., Ladle, A., et al. (2015). Wildlife camera trapping: a review and recommendations for linking surveys to ecological processes. Journal Applied Ecology 52(3): 675–685. Can, O. E., D’Cruze, N., Garshelis, D., Beecham, J. & MacDonald, D. W. (2014).

294

Resolving human–bear conflict: a global survey of countries, experts, and key factors. Conservation Letters 7(6): 510–513. Caughley, G. (1994). Directions in conservation biology. Journal of Animal Ecology 63(2): 215–244. Chapron, G., Kaczensky, P., Linnell, J. D. C., et al. (2014). Successful recovery of large carnivores in Europe’s human-dominated landscapes. Science 346(6216): 1517–1519. Cherry, S., Haroldson, M. A., Robison-Cox, J. & Schwartz, C.C. (2002). Estimating total human-caused mortality from reported mortality using data from radioinstrumented grizzly bears. Ursus 13: 175–184. CITES. (2000). Trade in bear specimens. Eleventh meeting of the Conference of the Parties Gigiri (Kenya), April 10–20, 2000. Gigiri, Kenya. Clapham, M., Miller, E. & Nguyen, M. (2018). Developing automated face recognition technology for non-invasive monitoring of brown bears (Ursus arctos). In: Skrbinšek, M. A. (Ed.), 26th International Conference on Bear Research and Management. University of Ljubljana, Abstract 237 (p. 22). Available from: https://lifewithbears.eu/book-ofabstracts/ Clark, T. W. (2002). The policy process. A practical guide for natural resource professionals. New Haven, CT: Yale University Press. Clark, T. W. & Wallace, R. L. (1999). The professional in endangered species conservation: an introduction to standpoint clarification. Endangered Species Update 16(1): 9–13. Convention on the Conservation of European Wildlife and Natural Habitats (2018). Recommendation No. 198 (2018) on the use of artificial feeding as a management tool of large carnivore populations and their prey, with a particular emphasis on the brown bear. 17 [document T-PVS(2018)7]. Available from https://rm.coe.int/list-ofdecisions-and-adopted-textes-of-the38th-meeting-of-the-bern-c/16808fde8d Cooch, E. G. & White, G. C. (2019). Program MARK; a gentle introduction. Available from www.phidot.org/software/mark/ docs/book/pdf/mark_book.zip Crête, M., Vandal, D., Rivest, L. P. & Potvin, F. (1991). Double counts in aerial surveys

to estimate polar bear numbers during the ice-free period. Arctic 44(4): 275–278. Crudge, B., Lees, C., Hunt, M., et al. (2019). Range-wide Conservation Action Plan for the sun bear, Helarctos malayanus, 2018–2028. Gland: IUCN SSC Bear Specialist Group/Free the Bears/ TRAFFIC. Curran, L. M., Trigg, S. N., McDonald, A. K., et al. (2004). Lowland forest loss in protected areas of Indonesian Borneo. Science 303: 1000–1003. Czetwertynski, S., Boyce, M. S. & Schmiegelow, F. K. (2007). Effects of hunting on demographic parameters of American black bears. Ursus 18: 1–17. Danilov, P. & Tirronen, K. (2011). Large predators in the Russian North-West. In: Beiträge zur Jagd- und Wildforschung 36 Proceedings, April 7–10, 2011 (pp. 19–25). Camp Reinsehlen, Germany. Debata, S., Swain, K. K., Sahu, H. K. & Palei, H. S. (2016). Human–sloth bear conflict in a human-dominated landscape of northern Odisha, India. Ursus 27: 90–98. Denali National Park. (2003). Bear–human conflict management plan, Denali National Park and Preserve. Center for Resources, Science, and Learning. US National Park Service. Derocher, A. E. (2012). Polar bears: A complete guide to their biology and behavior. Baltimore, MD: Johns Hopkins University Press. Derocher, A., Aars, J. A., Amstrup, S. C., et al. (2013). Rapid ecosystem change and polar bear conservation. Conservation Letters 6(5): 368–375. Dressel, S., Sandström, C. & Ericsson, G. (2015). A meta-analysis of studies on attitudes toward bears and wolves across Europe 1976–2012. Conservation Biology 29(2): 565–574. Dutta, T. J., Maldanado, E., Panwar, H. S. & Seidensticker, J. (2015). Genetic variation, structure, and gene flow in a sloth bear (Melursus ursinus) metapopulation in the Satpura-Maikal landscape of central India. PLoS ONE 10(5): e0123384. Dutton, A., Hepburn, J. C. & Macdonald, D. W. (2011). A stated preference investigation into the Chinese demand for farmed vs. wild bear bile. PLoS ONE 6: e21243. https://doi.org/10.1371/journal .pone.0021243 Eberhardt, L. L. (2002). A paradigm for population analysis of long-lived vertebrates. Ecology 83: 2841–2834.

Conservation and Management

EIA. (2015). Sin City: Illegal wildlife trade in Laos’ Golden Triangle Special Economic Zone. London, UK: EIA. Elfström, M., Davey, M. L., Zedrosser, A., et al. (2014). Do Scandinavian brown bears approach settlements to obtain high-quality food? Biological Conservation 178: 128–135. Ervin J. (2003). Protected area assessments in perspective. BioScience 53(9): 819–822. European Union. (1992). Council Directive 92/43/EEC of 21 May 1992 on the conservation of natural habitats and of wild fauna and flora. Available from https://eur-lex.europa.eu/legal-content/ EN/TXT/?uri=CELEX:31992L0043 Fahrig, L. (2003). Effects of habitat fragmentation on biodiversity. Annual Review of Ecology, Evolution, and Systematics 34: 487–515. Feng, Y., Siu, K., Wang, N., et al. (2009). Bear bile: dilemma of traditional medicinal use and animal protection. Journal of Ethnobiology and Ethnomedicine 5: 2. Fieberg, J. R., Shertzer, K W., Conn, P. B., Noyce, K. V. & Garshelis. D. (2010). Integrated population modeling of black bears in Minnesota: implications for monitoring and management. PLoS ONE 5(8): e12114. https://doi.org/10.1371/ jjournal.pone 0012114 Foley, K. E., Stengel, C. J. & Shepherd, C. R. (2011). Pills, powders, vials and flakes: The bear bile trade in Asia. Petaling Jaya, Selangor, Malaysia: TRAFFIC Southeast Asia. Fredriksson, G. (2005). Human–sun bear conflicts in east Kalimantan, Indonesian Borneo. Ursus 16: 130–137. Fredriksson, G., Wich, S. A. & Trisno. (2006). Frugivory in sun bears (Helarctos malayanus) is linked to El Niño-related fluctuations in fruiting phenology, East Kalimantan, Indonesia. Biological Journal of the Linnean Society 89(3): 489–508. Gangaas, K. E., Kaltenborn, B. P. & Andreassen, H. P. (2013). Geo-spatial aspects of acceptance of illegal hunting of large carnivores in Scandinavia. PLoS ONE 8(7): e68849. https://doi.org/10 .1371/journal.pone.0068849 Garshelis, D. (1997). The arrogance of ignorance – a commentary on the bear trade. International Bear News 6: 4–6. Garshelis, D. (2002). Misconceptions, ironies, and uncertainties regarding

trends in bear population. Ursus 13: 321–334. Garshelis, D. (2009). Family Ursidae (bears). In: Wilson, D.E. & Mittermeier, R.A. (Eds.), Handbook of the mammals of the world. Vol. 1. Carnivores (pp. 448–497). Barcelona: Lynx Editions. Garshelis, D. (2011). Andean bear density and abundance estimates – how reliable and useful are they? Ursus 22(1): 47–64. Garshelis, D. & Hristienko, H. (2006). State and provincial estimates of American black bear numbers versus assessments of population trend. Ursus 17(1): 1–7. Garshelis, D. & Steinmetz, R. (2017). Ursus thibetanus. IUCN Red List of Threatened Species. www.iucnredlist.org. Garshelis, D. & Visser, L. G. (1997). Enumerating megapopulations of wild bears using an integrated biomarker. Journal of Wildlife Management 61(2): 466–480. Garshelis, D., Joshi, A. R. & Smith, J. L. D. (1999a). Estimating density and relative abundance of sloth bears. Ursus 11: 87–98. Garshelis, D., Joshi, A. R., Smith, J. L. D. & Rice, C. G. (1999b). Sloth bear conservation action plan. In: Servheen, C., Herrero, S. & Peyton, B. (Eds.), Status survey and conservation action plan: Bears (chapter 12). Gland, Switzerland and Cambridge, UK: IUCN/SSC Bear and Polar Bear Specialist Groups. Garshelis, D., Gibeau, M. L. & Herrero, S. (2005). Grizzly bear demographics in and around Banff National Park and Kananaskis Country, Alberta. Journal of Wildlife Management 69(1): 277–297. Garshelis, D., Wang, H., Wang, D., et al. (2008). Do revised giant panda population estimates aid in their conservation? Ursus 19: 168–176. Gaston, K. J., Jackson, S. F., Cantu-Salazar, L. & Cruz-Pinon, G. (2008). The ecological performance of protected areas. Annual Review of Ecology, Evolution, and Systematics 39: 93–113. Geist, V. (1995). North American policies of wildlife conservation. In: Geist, V. & McTaggert-Cowan, I. (Eds.), Wildlife conservation policy (pp. 75–129). Calgary: Detselig Enterprises Limited. Geist, V., Mahoney, S. P. & Organ, J. F. (2001). Why hunting has defined the North American model of wildlife conservation. Transactions of the North

American Wildlife and Natural Resources Conference 66: 175–185. Gibeau, M. L. (2012). Of bears, chess, and checkers: moving away from pure science to solve problems. The Wildlife Professional 6: 62–64. Glacier National Park. (2010). Bear management plan, Glacier National Park. US National Park Service. Gong, J. E. & Harris, R. B. (2006). The status of bears in China: a country report. In: Japan Bear Network (Compiler), Understanding Asian bears to secure their future (pp. 96–101). Ibaraki, Japan: Japan Bear Network. Gong, M., Guan, T., Hou, M., Liu, G. & Zhou, T. (2017). Hopes and challenges for giant panda conservation under climate change in the Qinling Mountains of China. Ecology and Evolution 7: 596–605. Gosselin, J., Zedrosser, A., Swenson, J. E. & Peletier, F. (2014). The relative importance of direct and indirect effects of hunting mortality on the population dynamics of brown bears. Proceedings of the Royal Society B 282: 20141840. http://dx.doi.org/10.1098 .rsph.2014.1840. Gray, T. N. E., Lynam, A. J., Seng, T., et al. (2017). Wildlife-snaring crisis in Asian forests. Science 355: 255–256. Gunther, K. (Ed.). (2015). Grizzly bear recovery in the Yellowstone Ecosystem. Yellowstone Science 23(2): 4–46. Harris, R. B., Schwartz, C. C., Haroldson, M. A. & White, G. C. (2006). Trajectory of the Yellowstone grizzly bear population under alternative survival rates. In: Schwartz, C. C., Haroldson, M. A., White, G. C., et al. Temporal, spatial and environmental influences on the demographics of grizzly bears in the Greater Yellowstone Ecosystem (pp. 44–56). Wildlife Monographs 161. Harrison, R. D. (2011). Emptying the forest: hunting and the extirpation of wildlife from tropical nature reserves. BioScience 61: 919–924. Hashimoto, Y., Kaji, M., Sawada, H. & Takasuki, S. (2003). Five-year study on the autumn food habits of the Asiatic black bear in relation to nut production. Ecological Research 18(5): 485–482. Hatter, I. W., Mowat, G. & McLellan, B. N. (2018). Statistical population reconstruction to evaluate grizzly bear trends in British Columbia, Canada. Ursus 29(1): 1–12.

295

Conservation and Management

Heller, N. E. & Zaveleta, E. S. (2009). Biodiversity management in the face of climate change: a review of 11 years of recommendations. Biological Conservation 142: 14–32. Hertel, A. G., Steyaert, S. M. J. G., Zedrosser, A., et al. (2016). Bears and berries: species-species selection foraging on a patchily distributed food resource in a human-altered landscape. Behavioral Ecology and Sociobiology 70(6): 931–842. Hertel, A. G., Bischof, R., Langval, O., et al. (2018a). Berry production drives bottomup effects on body mass and reproductive success in an omnivore. Oikos 127: 197–207. Hertel, A. G., Langval, O., Swenson, J. E., Kindberg, J. & Zedrosser A. (2018b). Fluctuating mast production does not drive Scandinavian brown bear behavior. Journal of Wildlife Management 83(4): 657–668. https://doi.org/10.1002/jwmg .21619 Hilderbrand, G. V., Schwartz, C. C., Robbins, C. T., et al. (1999). The importance of meat, particularly salmon, to body size, population productivity, and conservation of North American brown bears. Canadian Journal of Zoology 77(1): 132–138. Hooker, M. J., Laufenberg, J. S., Ashley, A. K., Sylvest, J. T. & Chamberlain, M. J. (2015). Abundance and density estimation of the American black bear population in central Georgia. Journal of Wildlife Management 26(2): 107–115. Höss, M., Kohn, M., Pääbo, S., Knauer, F. & Schröder W. (1992). Excrement analysis by PCR. Nature 359: 199. Housty, W. G., Noson, A., Scoville, G. W., et al. (2014). Grizzly bear monitoring by the Heiltsuk people as a crucible for First Nation conservation practice. Ecology and Society 19(2): 70. Huber, D., Jakšić, Z., Frković, A., et al. (2008a). Brown bear management plan for the Republic of Croatia. Zagreb, Croatia: Ministry of Agriculture, Forestry and Water Management, and Ministry of Culture. Huber, D., Kusak, J., Majić-Skrbinšek, A., Majnarić, D. & Sindićić, M. (2008b). A multidimensional approach to managing the European brown bear in Croatia. Ursus 19(1): 22–32. Humm, J. M., McCowen, J. W., Scheick, B. K. & Clark, J. D. (2017). Spatially explicit population estimates for black bears based on cluster sampling. Journal of

296

Wildlife Management 81(7): 1187–1201. doi: 10.1002/jwmg.21294

Ursus arctos population. Wildlife Biology 17(2): 114–123. doi: 10.2981/10-100

IUCN. (2012). Resolutions and Recommendations World Conservation Congress, Jeju, Republic of Korea, September 6–15, 2012. IUCN SSC Bear Specialist Group.

Knott, E. J., Bunnefeld, N., Huber, D., et al. (2014). The potential impacts of changes in bear hunting policy for hunting organizations in Croatia. European Journal of Wildlife Research 60(1): 85–97.

John, D., Linnell, C., Trouwborst, A., et al. (2016). Border security fencing and wildlife: the end of the transboundary paradigm in Eurasia? PLoS Biology 14(6): e1002483.

Kojola, I., Hallikainen, V., Helle, T. & Swenson, J. E. (2018). Can only poorer European countries afford large carnivores? PLoS ONE 13(4): e0194711. https://doi.org/10.1371/journal .pone.0194711

Jones, N. A., Shaw, S., Ross, H., Witt, K. & Pinner, B. (2016). The study of human values in understanding and managing social ecological systems. Ecology and Society 21(1): 15. Joppa, L. N. & Pfaff, A. (2009). High and far: biases in the location of protected areas. PLoS ONE 4(12): e8273. https://doi.org/ 10.1371/journal.pone.0008273 Keay, J. A., Robbins, C. T. & Farley, S D. (2018). Characteristics of a naturally regulated grizzly bear population. Journal of Wildlife Management 82(4): 789–801. Kellert, S.R. (1994). Public attitudes toward bears and their conservation. International Conference on Bear Research and Management 9(1): 43–50. Kellert, S. R. (1996). The value of life: Biological diversity and human society. Washington, DC: Island Press. Kendall, K. C., Stetz, J. B., Roon, D. A., et al. (2008). Grizzly bear density in Glacier National Park, Montana. Journal of Wildlife Management 72(8): 1693–1705. Kendall, K. C., Stetz, J. B., Boulanger, J., et al. (2009). Demography and genetic structure of a recovering grizzly bear population. Journal of Wildlife Management 73(1): 3–16. Kendall, K. C., Graves, T. A., Royle, J. A., et al. (2019). Using bear rub data and spatial capture–recapture models to estimate trend in a brown bear population. Scientific Reports 9: 16804. Keyghobadi, N. (2007). The genetic implications of habitat fragmentation for animals. Canadian Journal of Zoology 85: 1049–1064. Kindberg, J., Ericsson, G. & Swenson, J. E. (2009). Monitoring rare or elusive large mammals using effort-corrected voluntary observers. Biological Conservation 142: 159–165. Kindberg, J., Swenson, J. E., Ericsson, G., et al. (2011). Estimating population size and trends of the Swedish brown bear

Kolenosky, G. B. (1986). The effects of hunting on an Ontario black bear population. Bears: Their Biology and Management 6: 45–55. Krishnasamy, K. & Shepherd, C. (2014). A review of the sun bear trade in Sarawak, Malaysia. TRAFFIC Bulletin 26: 37–40. Krofel, M., Spacapan, M. & Jerina, K. (2017). Winter sleep with room service: denning behaviour of brown bears with access to anthropogenic food. Journal of Zoology 302(1): 8–14. Kusak, J., Huber, D., Gomerčić, T., Schwaderer, G. & Gužvica, G. (2009). The permeability of a highway in Gorski Kotar (Croatia) for large mammals. European Journal of Wildlife Research 55: 7–21. Lamb, C. T., Mowat, G., Reid, A., et al. (2018). Effects of habitat quality and access management on the density of a recovering grizzly bear population. Journal of Applied Ecology 55(3): 1406–1417. Lasswell, H. D. (1971). A preview of policy sciences. New York, NY: American Elsevier. Leslie, P. H. (1945). On the use of matrices in certain population mathematics. Biometrika 33(3): 183–212. Li, S., Tan, H. Y., Wang, N., et al. (2016). Substitutes for bear bile for the treatment of liver diseases: research progress and future perspective. Evidence-based Complementary and Alternative Medicine 2016: 4305074. https://doi.org/10.1155/ 2016/4305074 Linnell, J. D. C., Swenson, J. E. & Andersen, R. (2001). Predators and people: conservation of large carnivores is possible at high human densities if management policy is favourable. Animal Conservation 4(4): 345–349. Linnell J., Salvatori, V. & Boitani, L. (2008). Guidelines for population level

Conservation and Management

management plans for large carnivores in Europe: A Large Carnivore Initiative for Europe report prepared for the European Commission (contract 070501/2005/ 424162/MAR/B2). Liu, F., McShea, W. J., Garshelis, D., et al. (2011). Human–wildlife conflicts influence attitudes but not necessarily behaviors: factors driving the poaching of bears in China. Biological Conservation 144: 538–547. doi:10.1016/j. biocon.2010.10.009. Liu, J., Hull, V., Yang, W., et al. (2016). Pandas and people: Coupling human and natural systems for sustainability. Oxford: Oxford University Press. Livingstone, E. & Shepherd, C. R. (2014). Bear farms in Lao PDR expand illegally and fail to conserve wild bears. Oryx 50: 1–9. Livingstone, E., Gomez, L. & Bouhuys, J. (2018). A review of bear farming and bear trade in Lao People’s Democratic Republic. Global Ecology and Conservation 13: e00380. https://doi.org/ 10.1016/j.gecco.2018.e00380 Loucks, C. J., Lu, Z., Dinerstein, E., et al. (2001). Giant pandas in a changing landscape. Science 294: 1465. Lunn, N. J., Servanty, S., Regehr, E. V., et al. (2016). Demography of an apex predator at the end of its range: impacts of changing sea ice on polar bears in Hudson Bay. Ecological Applications 26(5): 1302–1320. Mace, R. D., Waller, J. S., Manley, T. L., Lyon, L. J. & Zuring, H. (1996). Relationships among grizzly bears, roads, and habitat use in the Swan Mountains, Montana. Journal of Applied Ecology 33(6): 1395–1404. Mace, R. D., Carney, D. W., ChiltonRadandt, T., et al. (2012). Grizzly bear population vital rates and trend in the Northern Continental Divide Ecosystem, Montana. Journal of Wildlife Management 76(1): 119–128. Mahoney, S. P. (2004). The seven sisters: pillars of the North American wildlife conservation model. Bugle 21: 5. Malenfant, R. M., Davis, C. S., Cullingham, C. I. & Coltman, D. W. (2016). Circumpolar genetic structure and recent gene flow of polar bears: a reanalysis. PLoS ONE 11(3): e0148967. doi:10.1371/ journal.pone.0148967. Manfredo, M. J., Teel, T. L., Sullivan, L. & Dietsch, A. M. (2017). Values, trust, and

cultural backlash in conservation governance: the case of wildlife management in the United States. Biological Conservation 214: 303–311. Margono, B. A., Potapov, P. V., Turubanova, S., Stolle, F. & Hansen, M. C. (2014). Primary forest cover loss in Indonesia over 2000–2012. Nature Climate Change 4: 1–6. Mattson, D. J. & Merrill, T. (2002). Extirpations of grizzly bears in the contiguous United States, 1850–2000. Conservation Biology 16(4): 1123–1136. Maxwell, S. L., Fuller, R. A., Brooks, T. M. & Watson, J. E. M. (2016). Biodiversity: the ravages of guns, nets, and bulldozers. Nature News 536: 143–145. McLaughlin, G. P., Primm, S. & Rutherford, M. B. (2005). Participatory projects for coexistence: rebuilding civil society. In: Clark, T. W., Rutherford, M. B. & Casey, D. (Eds.), Coexisting with large carnivores: Lessons from Greater Yellowstone (pp. 177–210). Washington, DC: Island Press. McLellan, B. N. (2005). Sexually selected infanticide in grizzly bears: the effects of hunting on cub survival. Ursus 16: 141–156. McLellan, B. N. (2011). Implications of a high-energy and low-protein diet on the body composition, fitness, and competitive abilities of black (Ursus americanus) and grizzly (Ursus arctos) bears. Canadian Journal of Zoology 89(6): 546–558. McLellan, B. N. (2015). Some mechanisms underlying variation on vital rates on grizzly bear in an industrial landscape. Journal of Wildlife Management 79(5): 749–765. McLellan, B. N. & Hovey, F. W. (2001a). Habitats selected by grizzly bears in multiple use landscapes. Journal of Wildlife Management 65(1): 92–99. McLellan, B. N. & Hovey, F. W. (2001b). Natal dispersal of grizzly bears. Canadian Journal of Zoology 79: 838–844. McLellan, B. N., Hovey, F. W., Mace, R. D., et al. (1999). Rates and causes of grizzly bear mortality in the interior mountains of British Columbia, Alberta, Montana, Washington, and Idaho. Journal of Wildlife Management 63(3): 911–920. McLellan, B. N., Proctor, M. F., Huber, D. & Michel, S. (2016). Ursus arctos Brown bear. IUCN Red List of Threatened Species. www.iucnredlist.org.

McLellan, B. N., Mowat, G., Hamilton, T. & Hatter, I. (2017). Sustainability of the grizzly bear hunt in British Columbia, Canada. Journal of Wildlife Management 81(2): 218–229. McLellan, B. N., Mowat, G. & Lamb, C. T. (2018). Estimating unrecorded humancaused mortalities of grizzly bears in the Flathead Valley, British Columbia, Canada. PeerJ 6: e5781. Mealy, S. (1979 [revised 1986]). Guidelines for management involving grizzly bears in the Greater Yellowstone Area. US Forest Service and US National Park Service. Meijaard, E., Sheil, D., Nasi, R., et al. (2005). Life after logging: Reconciling wildlife conservation and production forestry in Indonesian Borneo. Bogor, Indonesia: CIFOR and UNESCO. Miettinen, J., Shi, C. & Liew, S. C. (2011). Deforestation rates in insular Southeast Asia between 2000 and 2010. Global Change Biology 17: 2261–2270. Mikle, N., Graves, T. A., Kovach, R., Kendall, K. C. & Macleod, A. (2016). Demographic mechanisms underpinning genetic assimilation of remnant groups of a large carnivore. Proceedings of the Royal Society B 283(1839): 20161467. Miller, S. D. (1990). Population management of bears in North America. International Conference on Bear Research and Management 8: 357–373. Miller, S. D. & Tutterrow, V. L. (1999). Characteristics of non-sport mortalities to brown and black bears and human injuries from bears in Alaska. Ursus 11: 239–252. Miller, S. D., White, G. C., Sellers, R. A., et al. (1997). Brown and black bear density estimation in Alaska using radiotelemetry and replicated mark–resight techniques. Wildlife Monographs 133. Miller, S. D., Sellers, R. A. & Keay, J. A. (2003). Effects of hunting on brown bear cub survival and litter size in Alaska. Ursus 14: 130–152. Miller, S. D., McLellan, B. N. & Derocher, A. E. (2013). Conservation and management of large carnivores in North America. In: Mahoney, S. P. (Ed.), Conservation and hunting in North America. The International Journal of Environmental Studies Monographs 70(3), 383–398. http://dx.doi.org/10.1080/00207233 .2013:806099. Miller, S. D., Schoen, J. W. & Schwartz, C. C. (2017). Trends in brown bear reduction

297

Conservation and Management

efforts in Alaska, 1980–2017. Ursus 28(2): 135–149. Mills, J. & Servheen, C. (1991). The Asian trade in bears and bear parts. Washington, DC: World Wildlife Fund. Millspaugh, J. J., Skalski, J. R., Townsend, R. L., et al. (2010). An evaluation of sex–age– kill (SAK) model performance. Journal of Wildlife Management 73(3): 442–451. Mukesh, S., Sharma, L. K., Charoo, S. A. & Sathyakumar, S. (2015). Conflict bear translocation: investigating population genetics and fate of bear translocation in Dachigam National Park, Jammu and Kashmir, India. PLoS ONE 10(8): e0132005. https://doi.org/10.1371/journal .pone.0132005 Müller, A. & Kaji, K. (2018). Wildlife policy and laws in East Asia. In: Leopold, B. D., Kessler, W. B. & Cummins, J. L. (Eds.) North American Wildlife Policy and Laws (pp. 445–479). Missoula, MT: Boone and Crockett Club. Murdiyarso, D., Dewi, S., Lawrence, D. & Seymour, F. (2011). Indonesia’s forest moratorium: a stepping stone to better forest governance? Working Paper 76. Bogor, Indonesia: CIFOR. Naughton-Treves L., Holland, M. B. & Brandon, K. (2005). The role of protected areas in conserving biodiversity and sustaining local livelihoods. Annual Review of Environment and Resources 30: 219–252. Newton, E. & Obbard, M. E. (2018). Can population reconstruction be used to estimate black bear abundance in Ontario? Ontario Ministry of Natural Resources and Forestry, Science and Research Branch, Peterborough, ON. Science & Research Information Report IR-12. www.researchgate.net/publication/ 323956404 Ngoprasert, D., Reed, D. H., Steinmetz, R. & Gale, G. A. (2012). Density estimation of Asian bears using photographic capture– recapture sampling based on chest marks. Ursus 23(2): 117–133. Ngoprasert, D., Gale, G. A., Vichitsoonthonkul, T. & Vaeokhaw, S. (2015). Population demographics and genetics of Asiatic black bear and sun bear in the World Heritage Dong Phayayen– Khao Yai Forest Complex. Bangkok, Thailand: King Mongkut's University of Technology. Nielsen, S. E., Herrero, S., Boyce, M. S., et al. (2004). Modeling the spatial distribution of human-caused grizzly bear mortalities

298

in the Central Rockies Ecosystem of Canada. Biological Conservation 120(1): 101–113. Nomura, F., Higashi, S., Ambu, L. & Mohamed, M. (2004). Notes on oil palm plantation use and seasonal spatial relationships of sun bears in Sabah, Malaysia. Ursus 15(2): 227–231. Obbard, M. E., Cattet, M. R. L., Moody, T., et al. (2006). Temporal trends in the body condition of southern Hudson Bay polar bears. Climate Change Research Information Note #3. Ontario: Applied Research and Development Branch. Obbard, M. E., Howe, E. J., Wall, L. L., et al. (2014). Relationships among food availability, harvest, and human–bear conflict at landscape scales in Ontario, Canada. Ursus 25(2): 98–110. Obbard, M. E., Stapleton, S., Middel, K.R., et al. (2015). Estimating the abundance of the Southern Hudson Bay polar bear subpopulation with aerial surveys. Polar Biology 38: 1713–1725.

Peacock, E., Sonsthagen, S. A., Obbard, M. L., et al. (2015). Circumpolar genetic structure and recent gene flow of polar bears: a reanalysis. PLoS ONE DOI:10.1371/journal.pone.011202. Peck, C. P., Van Manen, F. T., Costello, C., et al. (2017). Potential paths for malemediated gene flow to and from an isolated grizzly bear population. Ecosphere 8(10): e0196. doi.org/10.1002/ ecs2.1969. Pelton, M. R. (2003). Black bear, Ursus americanus. In: Feldhamer, G. A., Thompson, B. C. & Chapman, J. A. (Eds.), Wild mammals of North America (2nd edition, pp. 547–555). Baltimore, MD: The Johns Hopkins University Press. Penteriani, V., Melletti, M. & Delgado, M. M. (2010). Don’t feed the bears! Oryx 44: 169–170. Penteriani, V., Lopez-Bao, J. V., Bettega, C., et al. (2017). Consequences of brown bear tourism: a review. Biological Conservation 206: 169–180.

Oi, T. & Yamazaki, K. (Eds.). (2009). The status of Asiatic black bears in Japan. In: Understanding Asian bears to secure their future (pp. 122–133). Ibaraki, Japan: Japan Bear Network.

Penteriani, V., Delgado, M. M., Krofel, M., et al. (2018). Evolutionary and ecological traps for brown bears in human-modified landscapes. Mammal Review 48: 180–193.

Ordiz A., Rodríguez C., Naves C., et al. (2007). Distance-based criteria to identify minimum number of brown bear females with cubs in Europe. Ursus 18: 158–167.

Penteriani, V., Zarzo-Arias, A., NovoFernández, A, Bombieri, G. & LópezSánchez, C. A. (2019). Responses of an endangered brown bear population to climate change based on predictable food resource and shelter alterations. Global Change Biology 25: 1133–1151.

Ordiz, A., Kindberg, J., Sæbø, S., Swenson, J. E. & Støen, O.-G. (2014). Brown bear circadian behavior reveals human environmental encroachment. Biological Conservation 173: 1–9. Organ, J. F., Geist, V., Mahoney, S. P., et al. (2012). The North American model of wildlife conservation. The Wildlife Society Technical Review 12-04. Bethesda, MD: The Wildlife Society. Paetkau, D. (2003). An empirical exploration of data quality in DNA-based population inventories. Molecular Ecology 12(6): 1375–1387. Paetkau, D. Amstrup, S. C., Born, E. W., et al. (1999). Genetic structure of the world’s polar bear populations. Molecular Ecology 8: 1571–1584. Parques Nacionales Naturales de Colombia & Wildlife Conservation Society. (2018). Estrategia para la Conservación del Oso Andino en los Parques Nacionales Naturales de Colombia (2016–2031). Bogotá D.C.

Pereira, H. M., Leadley, P. W., Proença, V., et al. (2010). Scenarios for global biodiversity in the 21st century. Science 330(6010): 1496–1501. doi:10.1126/ science.1196624 Peterson, N. N. & Nelson, M. P. (2016). Why the North American model for wildlife conservation is problematic for modern wildlife management. Human Dimensions of Wildlife 22(1): 43–54. Pierson, J. C., Graves, T. A., Banks, S. C., Kendall, K. C. & Lindenmayer, D.B. (2018). Relationship between effective and demographic population size in continuously distributed populations. Evolutionary Applications 11(7): 1162–1175. Prestrud, P. & Stirling, I. (1994). The International Polar Bear Agreement and the current status of polar bear conservation. Aquatic Mammals 20: 113–124.

Conservation and Management

Pritchard, G. T. & Robbins, C. T. (1990). Digestive and metabolic efficiencies of grizzly and black bears. Canadian Journal of Zoology 68(8): 1645–1651. Proctor, M. F., McLellan, B. N., Strobeck, C. & Barclay, R. (2004). Gender specific dispersal distances of grizzly bears estimated by genetic analysis. Canadian Journal of Zoology 82: 1108–1118. Proctor, M. F., McLellan, B. N., Strobeck, C. & Barclay, R. (2005). Genetic analysis reveals demographic fragmentation of grizzly bears yielding vulnerably small populations. Proceedings of the Royal Society, London B 272: 2409–2416. Proctor, M. F., McLellan, B. N., Boulanger, J., et al. (2010). Ecological investigations of grizzly bears in Canada using DNA from hair, 1995–2005. Ursus 21(2): 169–188. Proctor, M. F., Paetkau, D., McLellan, B. N., et al. (2012). Population fragmentation and inter-ecosystem movements of grizzly bears in western Canada and the northern United States. Wildlife Monographs 180: 1–46. Proctor, M. F., Lamb, C. T. & MacHutchon, A. G. (2017). The grizzly dance of berries and bullets: The relationship between bottom up food resources, huckleberries, and top down mortality risk on grizzly bear population processes in southeast British Columbia. Kaslo, BC, Canada: Trans-border Grizzly Bear Project. Available from http:// transbordergrizzlybearproject.ca/ research/publications.html.

Proctor, M. F., McLellan, B. N., Stenhouse, G. B., et al. (2020b). The effects of roads and motorized human access on grizzly bear populations in British Columbia and Alberta, Canada. Ursus 2019(30e2): 16–39. Ratnayeke, S., van Manen, F. T., Pieris, R. & Praga, V. S. J. (2007). Landscape characteristics of sloth bear ranges in Sri Lanka. Ursus 18: 189–202. Ratnayeke, S., van Manen, F. T., Pieris, R. & Praga, V. S. J. (2014). Challenges of large carnivore conservation: sloth bear attacks in Sri Lanka. Human Ecology 42: 467–479. Redpath, S., Linnell, J. C. D., Festa-Bianchet, M., et al. (2017). Don’t forget to look down – collaborative approaches to predator conservation. Biological Reviews 92(4): 2157–2163. Regehr, E. V., Wilson, R. R., Rode, K. D., Runge, M. C. & Stern, H. (2017). Harvesting wildlife affected by climate change: a modelling and management approach for polar bears. Journal Applied Ecology 54: 1534–1543. Regehr, E. V., Hostetter, N. J., Wilson, R. R., et al. (2018). Integrated population modeling provides the first empirical estimates of vital rates and abundance for polar bears in the Chukchi Sea. Scientific Reports 8: 1–12. Reljić, S., Jerina, K., Nilsen, E. B., et al. (2018). Challenges for transboundary management of a European brown bear population. Global Ecology and Conservation, 16: e00488. https://doi.org/ 10.1016/j.gecco.2018.e00488.

Proctor, M. F., McLellan, B. N., Stenhouse, G. B., et al. (2018a). Resource roads and grizzly bears in British Columbia, and Alberta. Canadian Grizzly Bear Management Series, Resource Road Management. Kaslo, BC, Canada: Trans-border Grizzly Bear Project. Available from http:// transbordergrizzlybearproject.ca/ research/publications.html.

Rios-Uzeda, B. & Wallace, R. B. (2007). A preliminary density estimate for Andean bear using camera-trapping methods. Ursus 18(1): 124–128.

Proctor, M. F., Kasworm, W. F., Annis, K. M., et al. (2018b). Conservation of threatened Canada–USA trans-border grizzly bears linked to comprehensive conflict reduction. Human Wildlife Interactions 12(3): 348–372.

Robbins, C. T., Ben-David, M., Fortin, J. K. & Nelson, O. L. (2012). Maternal condition determines birth date and growth of newborn bear cubs. Journal of Mammalogy 93(2): 540–546.

Proctor, M. F., Kasworm, W. F., Teisberg, J. E., et al. (2020a). American black bear population fragmentation detected with pedigrees in the transborder Canada– United States region. Ursus 2020(31e1): 1–15.

Ripple, W. J., Miller, S. D., Schoen, J. W. & Rabinowitch, S. P. (2019). Large carnivores under assault in Alaska. PLoS Biology 17(1): e3000090. https://doi.org/ 10.1371/journal.pbio.3000090

Rolston, H. (1988). Human values and natural systems. Society and Natural Resources 1(1): 271–283. Royle, J., Chandler, A. B., Sollmann, R. & Gardner, B. (2014). Spatial capture– Recapture. Waltham, MA: Elsevier Academic Press.

Rutherford, M. B. & Clark, T. W. (2005). Coexisting with large carnivores: lessons from Greater Yellowstone. In: Clark, T. W., Rutherford, M. B. & Casey, D. (Eds.), Coexisting with large carnivores: Lessons from Greater Yellowstone (pp. 254–270). Washington, DC: Island Press. Rutherford, M. B., Gibeau, M. L., Clark, S. G. & Chamberlain, E. C. (2009). Interdisciplinary problem solving workshops for grizzly bear conservation in Banff National Park, Canada. Policy Sciences 42(2): 163–187. Ryan, C. W., Pack, J. C., Igo, W. K. & Billings, A. (2007). Influence of mast production on black bear non-hunting mortalities in West Virginia. Ursus 16(1): 46–53. Sæther, B. E., Engen, S., Swenson, J. E., Bakke, Ø. & Sandegren, F. (1998). Assessing the viability of Scandinavian brown bear, Ursus arctos, populations: the effects of uncertain parameter estimates. Oikos 83(2): 403–416. Sathyakumar, S, Kaul, R., Ashraf, N. V. K., Mookerjee, A. & Menon, V. (2012). National bear conservation and welfare action plan. India: Ministry of Environment and Forests, Wildlife Institute of India and Wildlife Trust of India. Schmidt, J. H., Rattenbury, K. L., Robison, H. L., Gorn, T. S. & Shults, B. S. (2017a). Using non-invasive mark–resight and sign occupancy surveys to monitor lowdensity brown bear populations across large landscapes. Biological Conservation 207: 47–54. Schmidt, J. H., Wilson, T. L., Thompson, W. L. & Reynolds, J. H. (2017b). Improving inference for aerial surveys of bears: the importance of assumptions and the cost of unnecessary complexity. Ecology and Evolution 7(13): 4812–4821. https://doi:10.1002/ece3.2912. Schwartz, C. C., Haroldson, M. A., Gunther, K. A. & Moody, D. (2002). Distribution of grizzly bears in the greater Yellowstone Ecosystem, 1990–2000. Ursus 13: 203–212. Schwartz, C. C., Miller, S. D. & Haroldson, M. A. (2003). Grizzly bear. In: Feldhamer, G. A., Thompson, B. C. & Chapman, J. A. (Eds.), Wild mammals of North America: Biology, management, and conservation (2nd edition, pp. 556–586). Baltimore, MD: The Johns Hopkins University Press. Schwartz, C. C., Haroldson, M. A., White, G. C., et al. (2006). Temporal, spatial, and

299

Conservation and Management

environmental influences on the demographics of grizzly bears in the Greater Yellowstone Ecosystem. Wildlife Monographs 161.

Servheen, C. & Mills, J. (1991). The trade in bears and bear parts. International Conference on Bear Research and Management 9: 161–167.

Schwartz, C. C., Haroldson, M. A. & White, G. C. (2010). Hazards affecting grizzly bear survival in the greater Yellowstone ecosystem. Journal of Wildlife Management 74(4): 654–667.

Servheen, C., Herrero, S. & Peyton, B. (Eds). (1999). Bears: Status survey and conservation action plan. Gland, Switzerland and Cambridge, UK: IUCN/ SSC Bear and Polar Bear Specialist Goups.

Schwartz, C. C., Fortin, J. K., Teisberg, J. E., et al. (2014). Body and diet composition of sympatric black and grizzly bears in the Greater Yellowstone Ecosystem. Journal of Wildlife Management 78(1): 68–78. Scotson, L., Vannachomchan, K. & Sharp, T. (2014). More valuable dead then deterred? Crop-raiding bears in Lao PDR. Wildlife Society Bulletin 38: 783–790. Scotson, L., Fredriksson, G., Ngoprasert, D., Wong, W. M. & Fieberg, J. (2017a). Projecting range-wide sun bear population trends using tree cover and camera-trap bycatch data. PLoS ONE 12(9): e0185336. https://doi.org/10.1371/ journal.pone.0185336. Scotson, L., Fredriksson, G., Augeri, D., et al. (2017b). Helarctos malayanus sun bear. IUCN Red List of Threatened Species. www.iucnredlist.org/species/9760/ 123798233. Scott, J. M., Goble, D. D., Wiens, J. A., et al. (2005) Recovery of imperiled species under the Endangered Species Act: the need for a new approach. Frontiers in Ecology and the Environment 3(7): 383–389. Scott, J. M., Goble, D. D., Haines, A. M., Weins, J. A. & Neel, M. C. (2010). Conservation reliant species and the future of conservation. Conservation Letters 3(2): 91–97. Seber, G. A. F. (1982). The estimation of animal abundance. London: Charles Griffin. Selva, N., Berezowska-Cnota, T. & ElgueroClaramunt, I. (2014). Unforeseen effects of supplementary feeding: ungulate baiting sites as hotspots for ground-nest predation. PLoS ONE 9(3): e90740. https://doi.org/10.1371/journal .pone.0090740 Servheen, C. (1990). The status and conservation of the bears of the world. International Association for Bear Research and Management, Monograph #2. Servheen, C. (1998). Conservation of small bear populations through strategic planning. Ursus 10: 67–73.

300

Servicio Nacional de Areas Naturales Protegidas por el Estado, Gobierno Regional del Cusco, Wildlife Conservation Society, INKATERRA. (2014). Estrategia para la Conservación del Oso Andino en el Santuario Histórico de Machupicchu y el Área de Conservación Regional Choquequirao. Lima, Perú. Sharma, S., Dutta, T., Maldonado, J. F., et al. (2013). Forest corridors maintain historic gene flow in a tiger metapopulation in the highlands of central India. Proceedings of the Royal Society 280: 20131506. Shepherd, C. R. & Shepherd, L. A. (2010). The poaching and trade of Malayan sun bears in peninsular Malaysia. Traffic Bulletin 23: 68. Singh, N., Sonone, S. & Dharaiya, N. (2018). Sloth bear attacks on humans in central India: implications for species conservation. Human–Wildlife Interactions 12(3): 338–347. https://doi .org/10.26077/2mgq-fs29 Skalski, J. R., Millspaugh, J. J. & Clawson, M. V. (2012). Comparison of statistical population reconstruction using full and pooled adult age-class data. PLoS ONE 7 (3):e33910. https://doi.org/10.1371/ journal.pone 0033910. Skrbinšek, T., Jelenčič, M., Luštrik, R., et al. (2018). Using laboratory robotics, highthroughput sequencing and sampling with volunteers for quick and costeffective large-scale genetic estimates of brown bear population size – transboundary case study in Slovenia. In: Skrbinšek, M. A. (Ed.), 26th International Conference on Bear Research and Management. University of Ljubljana, Abstract 211 (p. 76). Available from https://lifewithbears.eu/book-ofabstracts/

science to determine grizzly bear distribution and abundance in WestCentral, Alberta. Report prepared for Alberta Conservation Association and Alberta Environment and Parks. Available from https://friresearch.ca/ resource/using-scat-dna-and-citizenscience-determine-grizzly-beardistribution-and-abundance-west Soulé, M. E. (1985). What is conservation biology? A new synthetic discipline addresses the dynamics and problems of perturbed species, communities, and ecosystems. Bioscience 35: 727–734. Soulé, M. E. (1986). Conservation biology: The science of scarcity and diversity. Sunderland, MA: Sinauer Associates. Stapleton, S., Atkinson, S., Hedman, D. & Garshelis, D. (2014). Revisiting Western Hudson Bay using aerial surveys to update polar bear abundance in a sentinel population. Biological Conservation 170: 38–47. Stapleton, S., Peacock, E. & Garshelis, D. (2015). Aerial surveys suggest long-term stability in the seasonally ice-free Foxe Basin (Nunavut) polar bear population. Marine Mammal Science 32(1): 181–201. https://doi.org/10.1111/mms.12251. Steinmetz, R. & Garshelis, D. (2008). Distinguishing Asiatic black bears and sun bears by claw marks on climbed trees. Journal of Wildlife Management 72(3): 814–821. Stenhouse, G. B., Boulanger, J., Efford, M., et al. (2015). Estimates of grizzly bear population size and density for the 2014 Alberta Yellowhead Population Unit (BMA 3) and south Jasper National Park Inventory Project. Report prepared for Weyerhaeuser Ltd., West Fraser Mills Ltd., Alberta Environment and Parks, and Jasper National Park. Available from https://friresearch.ca/resource/estimatesgrizzly-bear-population-size-and-densityfinal-report. Stenset, N. E., Lutnæs, P. N., Bjarnadóttir, V., et al. (2016). Seasonal and annual variation in the diet of brown bears (Ursus arctos) in the boreal forest of southcentral Sweden. Wildlife Biology 22(3): 107–116.

Skuban, M., Findo, S. & Kajba, M. (2016). Human impacts on bear feeding habits and habitat selection in the Poľana Mountains, Slovakia. European Journal of Wildlife Research 62(3): 353–364.

Stetz, J. B., Kendall, K. C. & Servheen, C. (2010). Evaluation of bear rub surveys to monitor grizzly bear population trends. Journal of Wildlife Management 74(4): 860–870.

Sorenson, A., Milligan, S., Larsen, T., et al. (2017). Using scat DNA and citizen

Stetz, J. B., Kendall, K. C. & Macleod, A. C. (2014). Black bear density in Glacier

Conservation and Management

National Park, Montana. Wildlife Society Bulletin 38(1): 60–70.

Taberlet, P., Griffin, S., Goossens, B., et al. (1996). Reliable genotyping of samples with very low DNA quantities using PCR. Nucleic Acids Research 26: 3189–3194.

Steyaert, S. M. J. G., Zedrosser, A., Elfstrom, M., et al. (2016). Ecological implication from spatial patterns in human-caused brown bear mortality. Wildlife Biology 22(4): 144–152.

Tarsim. (2018). Agricultural insurance pool for Turkey. Available from https://web .tarsim.gov.tr/havuz/homePageEng

Stirling, I., Lunn, N. J. & Iacozza, J. (1999). Long-term trends in the population ecology of polar bears in western Hudson Bay in relation to climate change. Arctic 52(3): 294–306.

Taylor, M., DeMaster, D., Bunnell, F. L. & Schweinsburg, R. (1987). Modeling the sustainable harvest of female polar bears. Journal of Wildlife Management 51: 811–820.

Stirling, I., McDonald, T. L., Richardson, E. S., Regehr, E. V. & Amstrup, S. C. (2011). Polar bear population status in the northern Beaufort Sea, Canada 1971–2006. Ecological Applications 21(3): 859–876.

Tirronen, K. F., Panchenko, D. V. & Kusnecova, A. S. (2015). Brown bear (Ursus arctos L.) of the white sea of the Kola Peninsula. The Herald of Game Management 12(2): 125–136.

Stringham, S. F. (1980). Possible impacts of hunting on the grizzly/brown bear, a threatened species. International Conference on Bear Research and Management 5: 140–151. Swaisgood, R., Wang, D. & Wei, F. (2016). Ailuropoda melanoleuca. IUCN Red List of Threatened Species. wwwiucnredlistorg/details/712/0. Swaisgood, R., Wang, D. & Wei, F. (2018). Panda downlisted but not out of the woods. Conservation Letters 11: 1–9. Swenson, J. E. (2003). Implications of sexually selected infanticide for hunting of large carnivores. In: Festa-Bianchet, M. & Apollonio, M. (Eds.), Animal behaviour and wildlife conservation (pp. 171–190). Covelo, CA: Island Press. Swenson, J. E., Wabakken, P., Sandegren, F., et al. (1995). The near extinction and recovery of brown bears in Scandinavia in relation to the bear management policies of Norway and Sweden. Wildlife Biology 1(1): 11–25. Swenson, J.E., Sandegren, F., Soderberg, A., et al. (1997). Infanticide caused by hunting of male bears. Nature 386: 450–451. Swenson, J. E., Jansson, A., Riig, R. & Sandegren, F. (1999). Bears and ants: myrmecophagy by brown bears in central Scandinavia. Canadian Journal of Zoology 77(4): 551–561. Swenson, J. E., Schneider, M., Zedrosser, A., et al. (2017). Challenges of managing a European brown bear population; lessons from Sweden, 1943–2013. Wildlife Biology 2017(4): 1–13. wlb.00251 Taberlet, P. & Bouvet, J. (1992). Bear conservation genetics. Nature 358: 197.

Togawa, K. & Sakamoto, M. (2002). Bear markets: Japan, a summary of the findings of: “Japan’s illegal trade in bear products: A threat to bears worldwide”. Tokyo: Japan Wildlife Conservation Society. Tumbelaka, L. & Fredriksson, G. M. (2006). The status of sun bears in Indonesia (pp. 73–78). Ibaraki, Japan: Japan Bear Network. Tumendembergi, O., Proctor, M., Reynolds, H., et al. (2015). Gobi bear abundance and inter-oases movements, Gobi Desert, Mongolia. Ursus 26(2): 129–142. US Fish and Wildlife Service. (2016). National Survey of Fishing, Hunting, and Wildlife-Associated Recreation. Washington, DC: US Fish and Wildlife Service. van Manen, F. T., Haroldson, M. A., Bjornlie, D. D., et al. (2016). Density dependence, whitebark pine, and vital rates of grizzly bears. Journal of Wildlife Management 80(2): 300–313. van Manen, F. T., Haroldson, M. A. & Karabensh B. E. (Eds.). (2018). Yellowstone Grizzly Bear investigations: Annual report of the Interagency Grizzly Bear Study Team 2017. US Department of the Interior, US Geological Survey. Velez-Liendo, X. & Garcia-Rangel, S. (2017). Tremarctos ornatus. IUCN Red List of Threatened Species. www.iucnredlist.org. Viteri, M. P. (2007). Conservation genetics of Andean bears (Tremarctos ornatus) in northeastern Ecuador: molecular tools, genetic diversity and population size. PhD dissertation, University of Idaho, Moscow, Idaho, USA. Walsh, P., Reynolds, J., Collins, G., et al. (2010). Application of a double-observer aerial land-transect method to estimate

brown bear population density in southwestern Alaska. Journal Fish and Wildlife Management 1(1): 47–58. Wang, F., McShea, W. J., Wang, D., et al. (2014). Evaluating landscape options for corridor restoration between giant panda reserves. PLoS ONE 9: e105086. doi:10.1371/journal.pone.0105086. Wang, F., McShea, W. J., Wang, D. & Li, S. (2015). Shared resources between giant panda and sympatric wild and domestic mammals. Biological Conservation 186: 319–325. Wasser, S. K., Davenport, B., Ramage, E. R., et al. (2004). Scat detection dogs in wildlife research and management: application to grizzly and black bears in the Yellowhead Ecosystem, Alberta, Canada. Canadian Journal of Zoology 82(3): 475–492. doi:10.1139/Z04-020. Wei, F., Swaisgood, R., Hu, Y., et al. (2015). Progress in the ecology and conservation of giant pandas. Conservation Biology 29: 1497–1507. Wei, W., Swaisgood, R., Dai, Q., et al. (2018). Giant panda distributional and habitatuse shifts in a changing landscape. Conservation Letters 11: e12575. https:// doi.org/10.llll/conl.12575. Wilcove, D. S., Rothstein, D., Dubow, J., Phillips, A. & Losos, E. (1998). Quantifying threats to imperiled species in the United States. BioScience 48(8): 607–615. Willcox, D., Minh, N. D. T. & Gomez, L. (2016). An assessment of trade in bear bile and gall bladder in Viet Nam. Selangor, Malaysia: TRAFFIC, Petaling Jaya. Available from www.traffic.org/ publications/reports/bear-bile-and-gallbladder-trade-in-viet-nam/ Wilson, S. M., Neudecker, G. A. & Jonkel, J. J. (2014). Human–grizzly bear coexistence in the Blackfoot River Watershed, Montana: getting ahead of the conflict curve. In: Clark, S.G. & Rutherford, M. B. (Eds.), Large carnivore conservation: Integrating science and policy in the North American west (pp. 177–213). Chicago, IL: University of Chicago Press. Woodroffe, R. (2000). Predators and people: using human densities to interpret declines of large carnivores. Animal Conservation 3(2): 165–173. Wong, S. T., Servheen, C., Ambu, L. & Norhayati, A. (2005). Impacts of fruit production cycles on Malayan sun bears and bearded pigs in lowland tropical

301

Conservation and Management

forest of Sabah, Malaysian Borneo. Journal of Tropical Ecology 21(6): 627–639. Wong, W. M. & Linkie, M. (2013). Managing sun bears in a changing tropical landscape. Diversity and Distributions 19: 700–709. Wong, W. M., Leader-Williams, N. & Linke, M. (2013). Quantifying changes in sun bear distribution and their forest habitat in Sumatra. Animal Conservation 16(2): 216–223. Wong, W. M., Leader-Williams, N. & Linkie, M. (2015). Managing human–sun bear conflict in Sumatran agroforest systems. Human Ecology 43(2): 255–266. https:// doi.org/10.1007/s10745–015-9729-1 Woodroffe, R. & Ginsberg, J. R. (1998). Edge effects and the extinction of populations

302

inside protected areas. Science 280(5372): 2126–2128. Woods, J. G., Paetkau, D., Lewis, D., et al. (1999). Genetic tagging of free-ranging black and brown bears. Wildlife Society Bulletin 27(3): 616–627. World Animal Protection. (2017). Ending bear bile farming in South Korea. London, UK: World Animal Protection. WWF. (2018). Living Planet Report – 2018: Aiming higher. Grooten, M. & Almond, R. E. A. (Eds.). Gland, Switzerland: WWF. Yoganand, K., Rice, C. G., Johnsingh, A. J. T. & Seidensticker, J. (2006). Is the sloth bear in India secure? A preliminary report on distribution, threats and conservation requirements. Journal of the Bombay Natural History Society 103(2–3): 172–181.

Zedrosser, A., Steyaert, S., Swenson, J. E. & Gossow, H. (2011). Brown bear conservation and the ghost of persecution past. Biological Conservation 144(9): 2163–2170. Zhan, X., Ling, M., Zhang, Z., et al. (2006). Molecular censusing doubles giant panda population estimate in a key nature reserve. Current Biology 16(12): R451–R452 Zhan, X., Tao, Y., Li, M., et al. (2009). Accurate population size estimates are vital parameters for conserving the giant panda. Ursus 20(1): 56–62. Ziegltrum, G. J. (2004). Efficacy of black bear supplemental feeding to reduce conifer damage in western Washington. Journal of Wildlife Management. 68(3): 470–474.

Chapter

21

How Is Climate Change Affecting Polar Bears and Giant Pandas? Melissa Songer, Todd C. Atwood, David C. Douglas, Qiongyu Huang, Renqiang Li, Nicholas W. Pilfold, Ming Xu, and George M. Durner

Introduction Anthropogenic greenhouse gas emissions are the primary cause of climate change (Rosenzweig et al. 2008; Kokic et al. 2014) and an estimated increase of 3.7–4.8°C is predicted by the year 2100 if emissions continue at current levels (IPCC 2014). The main predictions for the physical impacts of climate change include changes in precipitation, humidity, and sea level, with higher temperatures over land and oceans leading to decreased snow cover, glaciers, and sea ice (Rosenzweig et al. 2007; Overland et al. 2019). While climate effects are spatially heterogeneous, varying in scale, magnitude, and type (IPCC 2018), their impacts are widely recognized as disruptive to ecological processes and a major threat to biodiversity globally. Changes to phenology (Lane et al. 2012), relationships and interactions within ecological communities (Post et al. 2013), and species ranges (Barnosky et al. 2003; Parmesan & Yohe 2003; Guralnick 2006) are predicted or already underway. Shifts in climatic niches can lead to trophic mismatches and phenological desynchronization that threaten species survival (Thackeray et al. 2016). Range shifts of terrestrial species to higher latitudes and altitudes have already been observed in the Northern Hemisphere, in addition to alterations in routes and timing of bird migrations in Europe, North America, and Australia, and changes in the composition of ocean communities from cold- to warm-adapted species (IPCC 2007). The Arctic is warming two to three times faster than temperate regions (IPCC 2018), and polar ecosystems are likely to be the first to undergo rapid alterations in climatic niches, with trophic shifts and impacts to biodiversity already occurring (Rosenzweig et al. 2007; Yurkowski et al. 2018). Polar bears (Ursus maritimus) and giant pandas (Ailuropoda melanoleuca) provide an interesting comparison study of the impact of climate change on bear species. While belonging to the same evolutionary clade, polar bears are the most recently divergent of the extant bear species, while giant pandas are the oldest (Yu et al. 2007). Both species are considered Vulnerable (Wiig et al. 2015; Swaisgood et al. 2017) and are conservation icons, polar bears in particular for climate change. Wildlife species with inflexible life histories and physiologies tightly regulated by the environment are most likely to experience declines in fitness associated with a changing climate. Examining how the environment structures

life histories and physiology of polar bears and giant pandas can inform our understanding of threats posed by climate change to both species’ persistence. Polar bears have a circumpolar distribution, living in an extreme and dynamic sea ice environment. As such, their physiology and life histories have evolved to cope with a habitat that changes on multiple temporal scales, from hourly to annually. Their physiology is adapted for a feast and famine regime, including the ability to rapidly gain energy-rich fat deposits during hyperphagia (April–October; Watts & Hansen 1987), while slowing metabolism to prolong fasting capacity during hypophagia (October–May; Pilfold et al. 2016). Polar bears also synchronize reproductive activities to resource seasonality (see also Chapter 14). Mating occurs April–June, and implantation occurs August–October, provided sufficient fat stores exist (Derocher et al. 1992). Cubs are born in November–January (Ramsay & Stirling 1988), and are nursed in a den until February–March (Yee et al. 2017), when they follow their mothers onto the sea ice to hunt vulnerable ringed seal (Pusa hispida) pups born during the period when polar bear hyperphagia peaks (Pilfold et al. 2012). The ringed seal whelping period is vital to polar bear cub survival, and changes in availability of ringed seal pups correlate to declines in polar bear reproductive output (Stirling 2002; see also Chapter 14). These changes have been linked to region-wide environmental changes (Rode et al. 2018), indicating tight bottom-up regulation of polar bear populations. Traits that make a species more vulnerable to disturbance include having a restricted geographic range, limited dispersal, low reproductive rates, and high dietary and habitat specialization (Peters & Darling 1985; Reid et al. 1989; McDonald & Brown 1992; IUCN 2008; Isaac et al. 2009). Giant pandas have a narrow geographic range, do not disperse over large distances, produce one cub every two to three years, and depend on bamboo for 99% of their diet (Reid et al. 1991; Zhao et al. 2013; Wei et al. 2015), suggesting vulnerability to change. Giant pandas live in temperate montane forests of central China, and although bamboo is widely available, they select it in discrete patches based on accessibility and nutrient quality (Wei et al. 2015). Giant pandas shift feeding to favor bamboo leaves from June to December and culm and shoots from February to May (Hansen et al. 2010), and their seasonal

303

Conservation and Management

movements, in part, reflect this selection (Zhang et al. 2014). Mating occurs from February to April and birthing from July to September (Zhu et al. 2001; see Chapter 6 for details on the species biology). Giant panda ancestors were once carnivorous and, while evolution to a bamboo diet allowed them to exploit an abundant resource, their gastrointestinal system is inefficient at digesting highly fibrous, low-calorie bamboo. Their behavior and physiology have evolved to maintain a positive energy balance while feeding on a nutrient-poor resource (Zhu et al. 2011). Despite morphological adaptations to aid in bamboo consumption (e.g. pseudo-thumb, mandible and jaw modifications, microflora in gut), giant pandas must spend much of their time feeding (Schaller et al. 1985) and their population dynamics are sensitive to stochastic changes in bamboo supply (Carter et al. 1999). Past bamboo die-offs due to flowering events have led to significant mortality (Reid et al. 1989), suggesting giant panda population dynamics are also regulated by bottom-up processes. While polar bears and giant pandas are arguably the most distant of the bear species with regard to life histories and behavior, both are likely to be significantly impacted by the broad-scale changes to their environment that are predicted to result from climate change. Herein, we review the conservation status of both species and their habitats, and present current and predicted evidence of the impacts of a changing climate on polar bear and giant panda survival.

Polar Bears Current Status of Polar Bears and Their Habitat Polar bears are dependent on sea ice for nearly all aspects of their life history (see Chapter 14). Climatic warming has caused significant declines in sea ice, the primary habitat of

polar bears. In 2018, Arctic sea ice extent maximum (14.48 million km2) and minimum (4.59 million km2) were estimated to be 7.4% and 26.2%, respectively, below the mean maximum and minimum extents for 1981–2010 (https://nsidc.org). Since 1979, annual maximum and minimum Arctic sea ice extent decreased 47,000 (3.1% decade–1) and 75,000 (10.4% decade–1) km2 year–1, respectively (Figure 21.1). Rates of sea ice loss have varied throughout the ranges of the estimated 20,000–25,000 wild polar bears in 19 recognized subpopulations (Figure 21.2; http://pbsg.npolar.no/), with ice-covered days declining by seven to 19 days decade–1 since 1979, depending on the subpopulation (Stern & Laidre 2016). Polar bear habitat preferences are driven by the composition and distribution of sea ice relative to the annual timing of critical life history requirements. Critical habitat includes medium- to high-concentration sea ice near edges, frequently near land, and over continental shelves or shallow waters (Durner et al. 2009). In the polar basin (where six subpopulations range), optimal habitat declined annually from 1% to 14% between 1985–1995 and 1996–2006 (Durner et al. 2009), and continued to 2015 in some regions (Lone et al. 2018). Despite losses in optimal habitat, the preference by polar bears for particular sea ice types and its distribution (i.e. concentration, proximity of ice edges, shallow seas) has remained relatively unchanged (Laidre et al. 2015; Wilson et al. 2016). Habitat loss, however, has impacted polar bears in several ways, including behavioral changes (Cherry et al. 2013; Rode et al. 2015b; Atwood et al. 2016a), diminished body condition, reduced survival, and numerical declines for some subpopulations (Regehr et al. 2007; Rode et al. 2010; Bromaghin et al. 2015; Lunn et al. 2016; Obbard et al. 2016, 2018). Future declines in sea ice will increase energetic costs to polar bears using the remaining ice due to ice fragmentation, increased swimming between floes, and increased movement to Figure 21.1 Median Arctic sea ice extent, comparing 1980–1989 to 2010–2017, for March (A) and September (B). Sea ice data source: National Snow and Ice Data Center, https://nsidc.org/.

304

Effects of Climate Change Figure 21.2 Polar bear range across the circumpolar Arctic showing the 19 subpopulations recognized by the International Union for the Conservation of Nature’s Polar Bear Specialist Group and associated ecoregions as defined by Amstrup et al. (2008). The characteristics used to define ecoregions are described in Chapter 14. (A black and white version of this figure will appear in some formats. For the color version, please refer to the plate section.)

compensate for faster ice drift (Durner et al. 2017). Hence, subpopulations currently showing no negative effects may be adversely impacted should habitat loss continue (Stirling et al. 2011). The status of some polar bear subpopulations may be a response to biological productivity rather than ice declines. The Davis Strait subpopulation increased from 1971 to 2007, despite deteriorating sea ice, due to an increasing prey base (Peacock et al. 2013). Low reproduction in Davis Strait and body condition declines in Baffin Bay bears, however, could be due to combined effects of high population density and sea ice declines (Rode et al. 2012; Peacock et al. 2013). Reproduction and body condition of Chukchi Sea polar bears remained stable between 1986–1994 and 2008–2011, despite an annual 44-day increase in ice-free days, because of high biological productivity in that region (Rode et al. 2014). This likely explains the high annual survival and abundance of Chukchi Sea polar bears (Regehr et al. 2018). While studies demonstrate that biological productivity may lessen the impacts of sea ice loss on polar bears in some regions, those mitigating effects will be moot if current rates of sea ice loss continue through the twenty-first century.

Future Impacts of Climate Change on Polar Bears and Their Habitat The greatest twenty-first-century threat to polar bears is anthropogenic greenhouse gas (GHG)-induced loss of sea ice (Atwood et al. 2016b) as this threatens extirpation of many subpopulations by 2100 (Amstrup et al. 2008). Climate models show Arctic surface air temperatures surpassing 1986–2005

levels by 2.2°C (mitigated GHG) to 8.3°C (unabated GHG) at century’s end, with sea ice reductions of 8–34% in February and 43–94% in September (Collins et al. 2013). Ninety percent of unabated GHG scenarios project a September ice-free Arctic Ocean for five consecutive years before 2100 (Collins et al. 2013). If GHG emissions continue unabated, in any year beyond 2090, the Arctic Ocean could be ice-free for 5 consecutive months (Douglas & Atwood 2017). Ice loss will impact some polar bear subpopulations more than others. For subpopulations in the Convergent Ice (CE) and Divergent Ice (DE) ecoregions (Figure 21.2; Amstrup et al. 2008; Castro de la Guardia et al. 2013), Durner et al. (2009) predicted decreases in optimal habitat between 0.9% and 6.5% decade–1 by 2099 compared to 1985–1995 based on moderately abated GHG scenarios. The annual cumulative extent of optimal habitat was projected to decline by 32% (from 1.5 million to 1.0 million km2) in the CE and DE by 2100. Castro de la Guardia et al. (2013) used a moderately abated GHG scenario to predict increased duration between annual break-up and freeze-up for sea ice habitat occupied by the Western Hudson Bay (WH) subpopulation (the period during which bears are forced ashore). Using 180 days on land as the threshold at which polar bear reproduction and survival would be compromised (Molnár et al. 2010, 2011), Castro de la Guardia et al. (2013) projected that this critical state would occur in 35 years between 2051 and 2100 and would jeopardize the persistence of WH polar bears. Unabated GHG emission scenarios had even more dire outcomes, with 41 years between 2051 and 2100 exceeding the critical threshold of 180 days. Amstrup et al. (2008) projected that diminishing sea ice will reduce numbers of animals in polar bear subpopulations

305

Conservation and Management

within all ecoregions by 2100. In the near term (i.e. within 25 years) subpopulations in the CE and the Canadian Arctic Archipelago Ecoregion (AE; Figure 21.2) had a relatively high probability of either remaining unchanged or increasing. Amstrup et al. (2008) suggested that the AE could provide refugium for polar bears in the late twenty-first century. However, this optimistic outlook contrasts with model projections based on aggressive emission scenarios that suggest the AE could change from a multiyear to annual ice environment, be ice-free for 2–5 months annually, and therefore unable to support polar bear populations by 2100 (Hamilton et al. 2014). The Seasonal Ice Ecoregion (SE) and DE subpopulations have a high probability of being reduced or even extirpated in the near term (Amstrup et al. 2008). Beyond 50 years, extirpation was the dominant outcome for the SE and DE subpopulations, and the dominant outcome for the CE subpopulation after 75 years.

Conservation Implications and Mitigation Strategies for Polar Bears Projections suggest that the circumpolar abundance of polar bears is likely to decline by 30–66% by mid-century if sea ice habitat loss continues unabated (Amstrup et al. 2008; Atwood et al. 2016b; Regehr et al. 2016). However, short-term (i.e. 10–20 years) responses to environmental changes may be mediated by a variety of factors, including regional differences in ecosystem productivity, anthropogenic stressors, and transient habitat improvements (Peacock et al. 2013; Rode et al. 2014; Jenssen et al. 2015). Because the long-term persistence of sea ice and polar bears are inextricably linked, actions that stop global warming and minimize sea ice loss must be a priority, but should not be the

sole focus, of polar bear conservation efforts. Diminishing sea ice is enabling greater human access to the Arctic, facilitating expanded oil, gas, and mineral extractions (Gautier et al. 2009), establishing new shipping lanes (Smith & Stephenson 2013), and providing more recreational opportunities (Rode et al. 2018). Noise associated with industrial development and recreation could cause polar bears to avoid impacted areas and lead to changes in the distribution and productivity of prey (Kelly et al. 1988). Increasing human activities could increase exposures to pollutants and emerging pathogens (Nuijten et al. 2016; Atwood et al. 2017a), as well as increase the risk for detrimental human–polar bear interactions (Atwood et al. 2017b). While such near-term stressors are inherently more tractable to mitigate than sea ice loss, they are likely to have far less direct influence on long-term population outcomes (Atwood et al. 2016b). Nevertheless, any reduction in shortterm population declines may prove crucial for maintaining viable polar bear subpopulations until such time that sea ice loss has ended. The best long-term conservation action for minimizing reductions in the number of polar bears and the amount of sea ice habitat upon which they rely would be to hold the global mean temperature to 2°C above pre-industrial levels. With prompt, aggressive, and sustained mitigation of GHG emissions, most climate models project that Earth’s average temperature would not rise more than 2°C and Arctic sea ice would persist all summer (Allen & Stocker 2014; Figure 21.3). Reductions to keep warming below 2°C would require immediate action because peak warmth from CO2 (a prevalent and long-lived GHG) emissions occurs a decade or more after release (Ricke & Caldeira 2014), and sea ice will require even more time to stabilize (Amstrup et al. 2010). Timely

Figure 21.3 Median monthly Arctic sea ice extent projections for the last decade of the twenty-first century. Medians were derived from projections by 13 general circulation models in the Coupled Model Intercomparison Project Phase 5 (CMIP5) when forced with the “worst case” (representative concentration pathways of greenhouse-related emissions or RCP 8.5; top row), “best case” (RCP 2.6; bottom row), and “intermediate case” (RCP 4.5; middle row) of anthropogenic greenhouse gas emission scenarios. Maps view the North Pole (center) and show sea ice (white), ocean (black), and land (gray). Figure from Douglas and Atwood (2017).

306

Effects of Climate Change

intervention in the current global warming trajectory, along with focused management of near-term stressors, will likely assure the persistence of polar bears through the twenty-first century (Amstrup et al. 2010).

Giant Pandas Current Status of Giant Pandas and Their Habitat Giant pandas were once distributed across the Yellow, Yangtze, and Pearl river basins in southeastern China, from Zhoukoudian in Beijing in the north to neighboring Vietnam, Thailand, and Myanmar in the south. Late Pleistocene climate changes and millennia of human development have reduced their geographic distribution to 100 individuals. Those larger populations are found in the northern half of the range – in the central Qinling, central Minshan, and northcentral Qionglai mountains. In the Minshan Mountains, 22 subpopulations have